1: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2: % %
3: % Site dilution of quantum spins in the honeycomb lattice %
4: % %
5: % Version date: 2005/08/04 %
6: % %
7: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
8:
9: \documentclass[twocolumn,prb,showpacs,preprintnumbers,amsmath,amssymb]{revtex4}
10: \usepackage{color}
11: \usepackage{graphicx}
12: \usepackage{amssymb}
13:
14: \makeatletter
15:
16: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%% LyX specific LaTeX commands.
17: %% Bold symbol macro for standard LaTeX users
18: \providecommand{\boldsymbol}[1]{\mbox{\boldmath $#1$}}
19:
20:
21: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%% User specified LaTeX commands.
22:
23: % Some other (several out of many) possibilities
24: %\documentclass[preprint,aps]{revtex4}
25: %\documentclass[preprint,aps,draft]{revtex4}
26: %\documentclass[prb]{revtex4}% Physical Review B
27: \usepackage{epsfig}
28: \usepackage{graphicx}% Include figure files
29: \usepackage{dcolumn}% Align table columns on decimal point
30: \usepackage{bm}% bold math
31: \usepackage{graphicx,color}
32:
33: % Just here to provide strikethrough capability
34: \usepackage{ulem}
35: \normalem
36: \newcommand{\kfix}[2]{\sout{#1} \textcolor{red}{#2}}
37: \newcommand{\kdel}[1]{\sout{#1}}
38: \newcommand{\kadd}[1]{\textcolor{red}{#1}}
39: \newcommand{\kcomment}[1]{\textcolor{blue}{#1}}
40:
41: \DeclareMathOperator{\diag}{diag}
42:
43: \makeatother
44: \begin{document}
45:
46: \title{Site dilution of quantum spins in the honeycomb lattice}
47:
48:
49: \author{Eduardo V. Castro,$^{1,2}$ N. M. R. Peres,$^{1,3,4}$
50: K. S. D. Beach$^{1}$, and Anders W. Sandvik$^{1}$}
51:
52:
53: \affiliation{$^{1}$Department of Physics, Boston University, 590 Commonwealth
54: Avenue, Boston, MA 02215, USA}
55:
56:
57: \affiliation{$^{2}$ CFP and Departamento de F\'{\i}sica, Faculdade de
58: Ci\^encias,
59: Universidade do Porto, P-4169-007 Porto, Portugal}
60:
61:
62: \affiliation{$^{3}$Max-Planck-Institut f\"ur Physik komplexer Systeme,
63: N\"othnitzer Str.\ 38, 01187 Dresden, Germany}
64:
65:
66: \affiliation{$^{4}$Center of Physics and Departamento de F\'{\i}sica,
67: Universidade do Minho, P-4710-057, Braga, Portugal}
68:
69:
70: \date{\today{}}
71:
72: \begin{abstract}
73: We discuss the effect of site dilution on both the magnetization and
74: the density of states of quantum spins in the honeycomb lattice, described
75: by the antiferromagnetic Heisenberg spin-$S$ model. Since the disorder
76: introduced by the dilution process breaks translational invariance,
77: the model has to be solved in real space. For this purpose a real-space
78: Bogoliubov-Valatin transformation is used. In this work we show that for
79: the $S>1/2$ the system can be analyzed in terms of linear spin wave theory,
80: in the sense that for all dilution concentrations the assumptions of validity
81: for the theory hold. For spin $S=1/2$, however, the linear spin wave
82: approximation breaks down. In this case, we have studied the effect of dilution
83: on the staggered magnetization using the Stochastic Series Expansion Monte
84: Carlo method. Two main results are to be stressed from the Monte Carlo
85: method: (i) a better value for the staggered magnetization of the
86: undiluted system, $m_{\rm av}(L\rightarrow\infty)=0.2677(6)$, relatively to
87: the only result available to date in the literature, and based on Trotter
88: error extrapolations; (ii) a finite value of the staggered magnetization
89: of the percolating cluster at the classical percolation threshold,
90: showing that there is no quantum critical
91: transition driven by dilution in the Heisenberg model. In the solution of the
92: problem using linear the spin wave method we pay special attention to the
93: presence of zero energy modes. We show, for a finite-size system (in a
94: bipartite lattice), that if the two sub-lattices are evenly diluted the system
95: always has two zero energy modes, which play the role of Goldstone boson modes
96: for a diluted lattice, having no translation symmetry but supporting
97: long range magnetic order. We also discuss the case when the two sub-lattices
98: are not evenly diluted. In this case, for finite size lattices,
99: the Goldstone modes are not a well defined concept, and special
100: care is needed in taking them into account
101: in order for sensible physical results can be obtained. Using a
102: combination of linear spin wave analysis and the recursion method
103: we were able to obtain the thermodynamic limit behavior of the density of
104: states for both the square and the honeycomb lattices. We have used both
105: the staggered magnetization and the density of states to analyze neutron
106: scattering experiments (determining the effect of dilution on the system's
107: magnetic moment) and N\'eel temperature measurements on quasi-two-dimensional
108: honeycomb systems. Our results are in quantitative agreement with experimental
109: results on Mn$_{p}$Zn$_{1-p}$PS$_{3}$ (a diluted $S=5/2$ system) and on the
110: Ba(Ni$_{p}$Mg$_{1-p}$)$_{2}$V$_{2}$O$_{8}$ (a diluted $S=1$ system).
111: Our work should stimulate further experimental research in Heisenberg
112: diluted-honeycomb systems.
113: \end{abstract}
114:
115:
116: \pacs{75.10.Jm, 75.50.Ee, 75.30.Ds, 75.40.Mg}
117:
118: \maketitle
119:
120:
121: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
122: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
123: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
124: \section{Introduction}
125:
126: The study of dilution and its effect on the magnetic properties of
127: antiferromagnetic materials is a central problem in modern condensed
128: matter theory.\cite{VMG+02,GSK+00,CCN01,MNC04,San02}
129: For the square lattice, a number of important experimental and theoretical
130: results have been reported.\cite{VMG+02,CCN01,MNC04,San02} For the honeycomb
131: lattice, there are some experimental results in the literature\cite{GSK+00}
132: already, but the corresponding theoretical understanding lags far behind.
133:
134: Insulating antiferromagnets are possible candidates for exhibiting
135: quantum critical points separating ordered from disordered phases.
136: The quantum corrections to the staggered magnetization of diluted
137: antiferromagnetic insulators became an important experimental and
138: theoretical topic after site dilution of La$_{2}$CuO$_{4}$ by non
139: magnetic impurities,\cite{VMG+02,CCR+91} such as Zn or Mg. Theoretical
140: studies interpreting the magnetic properties of these diluted systems
141: have been recently performed,\cite{MNC04,CCN02,San02,CCN01} showing
142: a good agreement between theory and experiment. A description of
143: the effect of dilution on the spin flop phase of La$_{2}$CuO$_{4}$
144: was attempted from the point of view of a simple mean field
145: theory,\cite{Peres03} with some qualitative agreement with experimental
146: results. In addition, the expectation of a magnetic quantum phase
147: transition driven by the interplay of dilution and quantum fluctuations
148: was shown not to occur in the antiferromagnetic Heisenberg model in a
149: square lattice.\cite{San02,MNC04}
150: In the undiluted case, on the other hand, it was shown that the
151: Heisenberg model itself is incapable of describing the high energy
152: part of the spin wave spectrum; a calculation starting from the
153: Hubbard model was shown to give the correct high energy behavior.
154: \cite{PA03,PA02,SSS02}
155:
156: The key role played by dimensionality in determining the behavior
157: of a system of quantum magnetic moments lends special importance
158: to the honeycomb lattice, which has the lowest possible co-ordination
159: in more than one dimension (see Fig.~\ref{cap:honey}). Realizations
160: of insulating antiferromagnets based on this lattice have already
161: been achieved both with and without magnetic dilution. Recently Spremo
162: \emph{et al}.~\cite{Spr+05} have studied the magnetic properties of
163: a metal-organic antiferromagnet on an undiluted but distorted honeycomb
164: lattice. The authors found good agreement between the theoretical
165: predictions obtained within the framework of a modified spin wave approach
166: and the experimental results for the magnetization as a function of
167: uniform external field and for the uniform zero-field susceptibility.
168:
169: Honeycomb layers are also found in transition-metal thiophosphates
170: MPS$_{3}$, where M is a first row transition metal. These compounds
171: are viewed as {}``perfect'' 2D magnetic systems because of the weak
172: van der Waals cohesion energy binding the layers. In each layer the
173: magnetic ions are arranged in a honeycomb lattice. Neutron diffraction
174: and magnetic susceptibility studies on MnPS$_{3}$, FePS$_{3}$, and
175: NiPS$_{3}$ antiferromagnets\cite{FBO+82,KSY83,JV92} ($S=5/2$, $S=2$,
176: and $S=1$, respectively) showed the existence of quite different
177: types of ordering among the different compounds. Whereas for FePS$_{3}$
178: and NiPS$_{3}$ the metal ions are coupled ferromagnetically to two
179: of the nearest neighbors and antiferromagnetically to the third, for
180: MnPS$_{3}$ all nearest neighbors interactions within a layer are
181: antiferromagnetic. In fact, it turns out that the simplest nearest-neighbor
182: antiferromagnetic Heisenberg model is a reasonable approximation for
183: the description of the magnetic properties in MnPS$_{3}$, although
184: the second- ($J_{2}$) and third-nearest-neighbor ($J_{3}$)
185: interactions---which are both also antiferromagnetic---are
186: not negligible for this compound ($J_{1}/J_{2}\sim10$ and
187: $J_{1}/J_{3} \sim4$).\cite{WRL+98} Substitution of magnetic
188: Mn$^{2+}$ ions by nonmagnetic Zn$^{2+}$ impurities showed that
189: long-range order (LRO) is lost at $p=0.46\pm0.03$ for
190: Mn$_{p}$Zn$_{1-p}$PS$_{3}$.\cite{CV96,GH98,GSK+00}
191: The fact that LRO is preserved for dilutions higher than the classical
192: percolation threshold for the honeycomb lattice,
193: $p_{\text{c}}\simeq0.7$, is attributed to the significance
194: of $J_{2}$ and $J_{3}$ in this compound.
195:
196: Recently, Rogado \emph{et al}.~\cite{RHL+02} have characterized the
197: magnetic properties of the $S=1$ honeycomb compound BaNi$_{2}$V$_{2}$O$_{8}$,
198: which can be described as a weakly anisotropic 2D Heisenberg
199: antiferromagnet.\cite{HvN+03} The magnetic Ni$^{2+}$ ions lie on
200: weakly coupled honeycomb layers, exhibiting antiferromagnetic
201: LRO close to 50 K. The doped compound
202: Ba(Ni$_{p}$Mg$_{1-p}$)$_{2}$V$_{2}$O$_{8}$ has a fraction
203: $1-p$ of the honeycomb layer sites substituted by Mg$^{2+}$---a
204: nonmagnetic ion. Magnetic susceptibility studies showed that the
205: N\'{e}el temperature is substantially reduced with increasing doping in
206: the range $0.84\leq p\leq1$. For $p=0.84$ the onset of antiferromagnetic
207: LRO occurs only at $T_{N}\simeq17$ K, a $T_{N}$ reduction of almost
208: 70\% relative to its undiluted value. It would be interesting
209: to know whether the suppression of antiferromagnetic LRO by
210: nonmagnetic impurities occurs at the classical percolation transition
211: $p_{\text{c}}\simeq0.7$, as predicted by our calculations (see below).
212:
213: In addition to these exciting experimental results, the theoretical
214: result of Mucciolo \emph{et al}.\ for the square lattice,\cite{MNC04}
215: where the vanishing of the staggered magnetization for the $S=1/2$
216: systems coincides with the classical percolation transition, opened
217: the naive expectation that for a 2D lattice with nonfrustrating nearest
218: neighbor interactions and a smaller number of neighbors, magnetic
219: quantum-phase transitions driven by the interplay of disorder and
220: quantum fluctuations could occur. The honeycomb lattice is the simplest
221: realization of such a lattice, for its coordination number is smaller than
222: that of the square lattice. On the other hand, large-scale quantum Monte
223: Carlo simulations of the square lattice have shown that the percolating
224: cluster actually has a robust long-range order,\cite{San02} in disagreement
225: with the spin wave calculation where this order vanishes very close
226: to the percolation point. Hence, spin wave theory is not reliable
227: at and close to the percolation point for $S=1/2$, and we expect
228: this break-down also for the honeycomb lattice at the percolation point.
229: This expectation is confirmed; we have performed quantum Monte Carlo
230: simulations that show only a rather modest reduction of the sublattice
231: magnetization of the percolating cluster, whereas there is no long-range
232: order in spin wave theory for $S=1/2$ in this case.
233:
234:
235: Experimental $S=1/2$ antiferromagnetic
236: systems with honeycomb lattice structure have already been reported
237: by Zhou \emph{et al}.~\cite{ZD91} in the A$_{2}$CuBr$_{4}$ salt,
238: where A is morpholinium (C$_{4}$H$_{10}$NO). Their data is well
239: described by a nearest-neighbor antiferromagnetic Heisenberg model,
240: but with two different couplings $J_{a}$ and $J_{b}$. To the best of our
241: knowledge, the dilution of this system has not yet been attempted.
242:
243: Motivated by the experimental results on diluted Mn$_{p}$Zn$_{1-p}$PS$_{3}$
244: and Ba(Ni$_{p}$Mg$_{1-p}$)$_{2}$V$_{2}$O$_{8}$ and by the possibility
245: of quantum phase transitions driven by the interplay of disorder and
246: quantum fluctuations, we study here the effect of site dilution on
247: the magnetic properties of the Heisenberg antiferromagnetic nearest-neighbor
248: model, for an arbitrary spin-$S$ value. Our study is performed both at zero
249: and finite temperatures. A first attempt to understand the effect of a
250: nonmagnetic defect on the properties of the $S=1/2$ 2D Heisenberg
251: antiferromagnet in the honeycomb lattice was made by
252: de Ch\^atel \emph{et al}.~\cite{dCC+04} In their mean-field approach, a single
253: impurity was introduced in clusters up to 12 spins. It is clear, however, that
254: their results can only be applied to systems with
255: dilutions up to $1-p=1/13$. Moreover, the random nature of defects
256: cannot be accounted for using their method.
257:
258: In this paper, we follow the general idea of the work of Mucciolo \emph{et
259: al}.,~\cite{MNC04} by using the linear spin wave approximation
260: in real space to compute different physical quantities. In addition
261: we use finite-size scaling to determine the magnetic moment
262: of the samples. We address the problem
263: of determining the density of states (DOS) of our system
264: using a different and more reliable method,
265: which gives the behavior of the DOS in the thermodynamic
266: limit. The paper is organized as follows: in Sect.~\ref{sec:hamilt}
267: we present the Hamiltonian and the formalism we use; in Sect.~\ref{sec:numeric}
268: we give the numerical details of our method; we present the results
269: on the staggered magnetization and on density of states as well as
270: on the cluster characterization in Sect.~\ref{sec:results}; finally,
271: in Sect.~\ref{sec:conclusions} we summarize our work and present some
272: concluding remarks.
273:
274:
275: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
276: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
277: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
278: \section{Model Hamiltonian and formalism}
279: \label{sec:hamilt}
280:
281: The Heisenberg Hamiltonian describing quantum spins in a site-diluted
282: honeycomb lattice is written as
283: \begin{equation}
284: H=J\sum_{i\in A,\bm\delta}\eta_{i}\eta_{i+\bm\delta}
285: \mathbf{S}_{i}^{a}\cdot\mathbf{S}_{i+\bm\delta}^{b}\,,
286: \label{eq:hamilt1}
287: \end{equation}
288: where $\mathbf{S}_{i}^a$ ($\mathbf{S}_{i}^b$) is the spin
289: operator on a site $i$ of sublattice $A$ ($B$). The notation $i+\bm\delta$
290: represents a nearest neighbor site of site $i$, connected to $i$
291: by the vector $\bm\delta$. There are three different $\bm\delta$
292: vectors given by
293: \begin{equation}
294: \bm\delta_{1}=\frac{c}{2}(1,\sqrt{3})\,,\hspace{0.25cm}
295: \bm\delta_{2}=\frac{c}{2}(1,-\sqrt{3})\,,\hspace{0.25cm}
296: \bm\delta_{3}=-c(1,0)\,,
297: \end{equation}
298: where $c$ is the hexagon side length. The $\eta_{i}$ variables
299: can have the values 0 or 1 depending on whether the site $i$ exists
300: or not.
301:
302: The usual spin wave approximation starts by assuming that LRO exists and,
303: in the case of antiferromagnetism, that the ground state is not substantially
304: different from the N\'{e}el state. The mathematical meaning of this similarity
305: is that the following inequalities should hold:
306: \begin{alignat}{2}
307: S-\bigl\langle S_{i}^{a,z}\bigr\rangle
308: \ll S & \quad & \text{for $i\in A$,}\label{eq:NsGA}\\
309: S+\bigl\langle S_{i}^{b,z}\bigr\rangle
310: \ll S & \quad & \text{for $i\in B$.} \label{eq:NsGB}
311: \end{alignat}
312: With these in mind we express the spin operators in terms of bosonic
313: creation and annihilation operators as introduced by Holstein and
314: Primakoff.\cite{HP40} Holstein-Primakoff transformation is defined
315: for sublattice $A$ as
316: \begin{equation} \label{eq:sopa}
317: \begin{split}
318: S_{i}^{a,z} & = S-a_{i}^{\dagger}a_{i}\,, \\
319: S_{i}^{a,+} & = \sqrt{2S}\sqrt{1-\frac{a_{i}^{\dagger}a_{i}}{2S}}\, a_{i}\,, \\
320: S_{i}^{a,-} & = \sqrt{2S}\, a_{i}^{\dagger}
321: \sqrt{1-\frac{a_{i}^{\dagger}a_ {i}}{2S}}\,.
322: \end{split}
323: \end{equation}
324: In sublattice $B$ the spin have $S_{z}=-S$ projection in the N\'{e}el
325: state. Since the bosons should describe excitations above the ground
326: state, and this has to be such that inequalities \eqref{eq:NsGA}
327: and \eqref{eq:NsGB} are verified, the $S^{b,z}$ operator needs to
328: be redefined as $S_{i}^{b,z}=-S+b_{i}^{\dagger}b_{i}$. Accordingly,
329: the $S_{i}^{b,+}$ operator must create bosons, and all the operators
330: in sublattice $B$ are defined as
331: \begin{equation} \label{eq:sopb}
332: \begin{split}
333: S_{i}^{b,z} & = -S+b_{i}^{\dagger}b_{i}\,, \\
334: S_{i}^{b,-} & = \sqrt{2S}\sqrt{1-\frac{b_{i}^{\dagger}b_{i}}{2S}}\, b_{i}\,, \\
335: S_{i}^{b,+} & = \sqrt{2S}\, b_{i}^{\dagger}
336: \sqrt{1-\frac{b_{i}^{\dagger}b_{i}}{2S}}\,.
337: \end{split}
338: \end{equation}
339: It is worth mentioning that inequalities \eqref{eq:NsGA} and \eqref{eq:NsGB}
340: can also be expressed in terms of the new bosonic operators $a$ and
341: $b$ as
342: \begin{alignat}{2}
343: \bigl\langle a_{i}^{\dagger}a_{i}\bigr\rangle
344: \ll S & \quad & \text{for $i\in A$,} \label{eq:ineqAb} \\
345: \bigl\langle b_{i}^{\dagger}b_{i}\bigr\rangle
346: \ll S & \quad & \text{for $i\in B$,} \label{eq:ineqBb}
347: \end{alignat}
348: from which the linear spin wave approximation follows straightforwardly
349: by expanding the square roots in Eqs.~\eqref{eq:sopa} and \eqref{eq:sopb}
350: in powers of $1/S$ and keeping only the zeroth order terms:
351: \begin{equation} \label{eq:Sapprox}
352: \begin{aligned}
353: S_{i}^{a,+} & \simeq \sqrt{2S}\, a_{i}\,, \\
354: S_{i}^{a,-} & \simeq \sqrt{2S}\, a_{i}^{\dagger}\,,
355: \end{aligned}
356: \quad
357: \begin{aligned}
358: S_{i}^{b,-} & \simeq \sqrt{2S}\, b_{i}\,, \\
359: S_{i}^{b,+} & \simeq \sqrt{2S}\, b_{i}^{\dagger}\,,
360: \end{aligned}
361: \end{equation}
362: Inserting the resultant approximation \eqref{eq:Sapprox} into
363: Eq.~\eqref{eq:hamilt1} produces the linear spin wave Hamiltonian,
364: which reads
365: \begin{multline} \label{eq:haf}
366: H = -Jh_{a}S(S+1)\sum_{i\in A,\bm\delta}\eta_{i}\eta_{i+\bm\delta} \\
367: + JS\sum_{i\in A,\bm\delta}\eta_{i}\eta_{i+\bm\delta}\Bigl[h_{a}
368: \bigl(a_{i}a_{i}^{\dagger}+b_{i+\bm\delta}^{\dagger}b_{i+\bm\delta}\bigr) \\
369: + a_{i}b_{i+\bm\delta}+b_{i+\bm\delta}^{\dagger}a_{i}^{\dagger}\Bigr]\,.
370: \end{multline}
371: Note that we have introduced a magnetic anisotropy $h_{a}$ in the
372: $S_{i}^{a,z}S_{i+\bm{\delta}}^{b,z}$ term.
373:
374: The linear spin wave Hamiltonian \eqref{eq:haf} can be seen as having
375: a classical part of the form
376: \begin{equation}
377: H_{\text{cl}}=-Jh_{a}S(S+1)\sum_{i\in A,\bm\delta}\eta_{i}\eta_{i+\bm\delta}\,,
378: \label{eq:Hcl}
379: \end{equation}
380: and a quantum fluctuating part, which can be written as
381: \begin{equation}
382: H_{\text{sw}}=(\{ a\},\{ b^{\dagger}\})\mathbf{D}
383: (\{ a\},\{ b^{\dagger}\})^{\dagger}\,,
384: \label{eq:Hsw}
385: \end{equation}
386: where $(\{ a\},\{ b^{\dagger}\})^{\dagger}$ is a column
387: vector containing all the boson operators and
388: \begin{equation} \label{eq:matrixD}
389: \mathbf{D}=\begin{pmatrix}
390: \mathbf{K}^{a} & \mathbf{\Delta}\\
391: \mathbf{\Delta}^{T} & \mathbf{K}^{b}
392: \end{pmatrix}
393: \end{equation}
394: is the so-called grand dynamical matrix. For a diluted lattice,
395: the number of sites in sublattice $A$ need not be the same as that
396: in sublattice $B$; therefore the dimensions of the blocks in
397: $\mathbf{D}$ are $N_{a}\times N_{a}$ for $\mathbf{K}^{a}$,
398: $N_{b}\times N_{b}$ for $\mathbf{K}^{b}$, $N_{a}\times N_{b}$ for
399: $\mathbf{\Delta}$ and $N_{b}\times N_{a}$ for $\mathbf{\Delta}^{T}$.
400: The corresponding matrix elements are
401: \begin{flalign}
402: \quad K_{ij}^{a}&=
403: Jh_{a}S\delta_{ij}\eta_{i}\sum_{\bm\delta}\eta_{i+\bm\delta}\,,
404: & \text{for $i\in A$,} \label{eq:Ka}\\
405: \quad K_{ij}^{b}&=
406: Jh_{a}S\delta_{ij}\eta_{i}\sum_{\bm\delta}\eta_{i+\bm\delta}\,,
407: & \text{for $i\in B$,} \label{eq:Kb}\\
408: \quad \Delta_{ij}&=JS\eta_{i}\eta_{j}\,,
409: & \text{for $i\in A$, $j\in i+\bm\delta$,} \\
410: \quad \Delta_{ij}^{T}&=JS\eta_{i}\eta_{j}\,,
411: & \text{for $i\in B$, $j\in i+\bm\delta$.} \label{eq:DeltaT}
412: \end{flalign}
413:
414: The diagonalization of the bosonic Hamiltonian amounts to find a transformation
415: $\mathbf{T}$ such that
416: \begin{equation}
417: (\mathbf{T}^{\dagger})^{-1}\mathbf{D}\mathbf{T}^{-1}=
418: \diag(\omega_{1},\ldots,\omega_{N_{a}},\omega_{N_{a}+1},\ldots,
419: \omega_{N_{a}+N_{b}})\,,
420: \label{eq:diag}
421: \end{equation}
422: where $\diag(\omega_{1},\ldots, \omega_{N_a+N_b})$ stands for a diagonal matrix
423: with elements $\omega_{1},\ldots, \omega_{N_a+N_b}$ in its diagonal,
424: $N_{a}+N_{b}$ in number.
425: In this case all the eigenvalues $\omega_{1},\ldots,\omega_{N_{a}},
426: \omega_{N_{a}+1},\ldots,\omega_{N_{a}+N_{b}}$ are positive.
427: The quasi-particles associated with those eigenvalues are obtained from
428: \begin{equation}
429: (\{\alpha\},\{\beta^{\dagger}\})^{\dagger}=
430: \mathbf{T}(\{ a\},\{ b^{\dagger}\})^{\dagger}\,.
431: \label{eq:rot}
432: \end{equation}
433:
434: In the undiluted case, it is well known that Eq.~\eqref{eq:Hsw} can
435: be diagonalized through a Bogoliubov-Valatin transformation in the
436: reciprocal space. For a subsequent analysis it is convenient to reproduce
437: here the results of the calculation for the undiluted honeycomb
438: lattice.\cite{PAB04} We first introduce the operators $a_{\mathbf{k}}$
439: and $b_{\mathbf{k}}$ defined as the inverse Fourier transforms of
440: $a_{i}$ and $b_{i}$,
441: \begin{equation}
442: a_{i}=\frac{1}{\sqrt{N_{a}}}
443: \sum_{\mathbf{k}}e^{-i\mathbf{k}\cdot\mathbf{r}_{i}}a_{\mathbf{k}}\,,
444: \hspace{0.21cm}
445: b_{i}=\frac{1}{\sqrt{N_{b}}}
446: \sum_{\mathbf{k}}e^{-i\mathbf{k}\cdot\mathbf{r}_{i}}b_{\mathbf{k}}\,,
447: \label{eq:akbk}
448: \end{equation}
449: where the $\mathbf{k}$ summation ranges over the first Brillouin
450: zone of either sublattice $A$ or $B$. (Do not confuse the site index $i$
451: and the complex imaginary unit also present in the Fourier
452: transform). The vector $\mathbf{r}_{i}$ is the position vector of
453: site $i$, and $N_{a}=N_{b}$ in the absence of dilution. Substituting
454: Eq.~\eqref{eq:akbk} into Hamiltonian~\eqref{eq:Hsw} gives us
455: $H_{\text{sw}}=\sum_{\mathbf{k}}H_{\mathbf{k}}$, with
456: \begin{multline} \label{eq:Hk}
457: H_{\mathbf{k}}=JSz\Bigl[h_{a}\bigl(a_{\mathbf{k}}a_{\mathbf{k}}^{\dagger}
458: + b_{-\mathbf{k}}^{\dagger}b_{-\mathbf{k}}\bigr)\\
459: +\phi_{\mathbf{k}}a_{\mathbf{k}}b_{-\mathbf{k}}
460: +\phi_{\mathbf{k}}^{*}b_{-\mathbf{k}}^{\dagger}a_{\mathbf{k}}^{\dagger}\Bigr],
461: \end{multline}
462: where $\phi_{\mathbf{k}}$ is defined as
463: \begin{equation}
464: \phi_{\mathbf{k}}=\frac{1}{z}\sum_{\bm\delta}e^{-i\mathbf{k}\cdot\bm\delta}\,.
465: \label{eq:phik}
466: \end{equation}
467: The diagonalized form of Hamiltonian \eqref{eq:Hk}, given by
468: \begin{equation} \label{eq:Hkdiag}
469: H_{\mathbf{k}}=\omega_{\mathbf{k}}
470: \bigl(1+\alpha_{\mathbf{k}}^{\dagger}\alpha_{\mathbf{k}}
471: +\beta_{\mathbf{k}}^{\dagger}\beta_{\mathbf{k}}\bigr)\,,
472: \end{equation}
473: with
474: \begin{equation}
475: \omega_{\mathbf{k}}=JSz\sqrt{h_{a}^{2}-\left|\phi_{\mathbf{k}}\right|^{2}},
476: \label{eq:Ek}
477: \end{equation}
478: can be easily obtained from the following Bogoliubov-Valatin transformation,
479: \begin{equation} \label{eq:BVT}
480: \begin{split}
481: \alpha_{\mathbf{k}} & =
482: u_{\mathbf{k}}a_{\mathbf{k}}+v_{\mathbf{k}}b_{-\mathbf{k}}^{\dagger}\,, \\
483: \beta_{\mathbf{k}} & =
484: v_{\mathbf{k}}a_{\mathbf{k}}^{\dagger}+u_{\mathbf{k}}b_{-\mathbf{k}}\,,
485: \end{split}
486: \end{equation}
487: with coefficients $u_{\mathbf{k}}$ and $v_{\mathbf{k}}$
488: given as functions of the parameters $h_{a}$ and $\phi_{\mathbf{k}}$.
489:
490: In the diluted case, translational symmetry is lost, and the solution
491: in the reciprocal space is as difficult as the one in real space. Let us start
492: by constructing an operator transformation of the Bogoliubov-Valatin
493: type in real space which can be used in the presence of dilution:
494: \begin{align}
495: \alpha_{n}&=\sum_{i=1}^{N_{a}}u_{ni}a_{i}+
496: \sum_{i=1}^{N_{b}}v_{ni}b_{i}^{\dagger}\,, \label{eq:alpha}\\
497: \beta_{n}&=\sum_{i=1}^{N_{a}}w_{ni}a_{i}^{\dagger}+
498: \sum_{i=1}^{N_{b}}x_{ni}b_{i}\,.\label{eq:beta}
499: \end{align}
500: This definition of $\alpha^{\dagger}$ and $\beta^{\dagger}$ excitations
501: gives us $N_{a}$ $\alpha$-type quasi-particles and $N_{b}$
502: $\beta$-type quasi-particles. Equations \eqref{eq:alpha}
503: and \eqref{eq:beta} define the transformation matrix \eqref{eq:rot},
504: \begin{equation} \label{eq:T}
505: \mathbf{T}=\begin{pmatrix}
506: \mathbf{U}^{*} & \mathbf{V}^{*}\\
507: \mathbf{W} & \mathbf{X}\end{pmatrix},
508: \end{equation}
509: where the $N_{a}\times N_{a}$ ($N_{b}\times N_{a}$) and $N_{a}\times N_{b}$
510: ($N_{b}\times N_{b}$) matrices $\mathbf{U}^{*}$ ($\mathbf{W}$)
511: and $\mathbf{V}^{*}$ ($\mathbf{X}$) contain the coefficients $u_{ni}^{*}$
512: ($w_{ni}$) and $v_{ni}^{*}$ ($x_{ni}$), respectively. Since the
513: quasi-particles must have a bosonic character, the quasi-particle operators
514: must satisfy the commutation relations
515: \begin{align}
516: \bigl[\alpha_{n},\alpha_{m}^{\dagger}\bigr]=
517: \bigl[\beta_{n},\beta_{m}^{\dagger}\bigr] & = \delta_{nm}\,,\\
518: \bigl[\alpha_{n},\beta_{m}\bigr]=
519: \bigl[\beta_{n}^{\dagger},\alpha_{m}^{\dagger}\bigr] & = 0\,,
520: \end{align}
521: which lead to the following constraints on the transformation coefficients:
522: \begin{align}
523: \sum_{i=1}^{N_{a}}u_{ni}u_{mi}^{\ast}-\sum_{i=1}^{N_{b}}v_{ni}v_{mi}^{\ast}
524: & = \delta_{nm}\,,\label{eq:ORalpha}\\
525: \sum_{i=1}^{N_{a}}w_{ni}w_{mi}^{\ast}-\sum_{i=1}^{N_{b}}x_{ni}x_{mi}^{\ast}
526: & = -\delta_{nm}\,,\label{eq:ORbeta}\\
527: \sum_{i=1}^{N_{a}}u_{ni}w_{mi}-\sum_{i=1}^{N_{b}}v_{ni}x_{mi}
528: & = 0\,,\label{eq:ORaphbet}\\
529: \sum_{i=1}^{N_{a}}w_{ni}^{\ast}u_{mi}^{\ast}-
530: \sum_{i=1}^{N_{b}}x_{ni}^{\ast}v_{mi}^{\ast} & = 0\,.
531: \label{eq:ORalphbetcc}
532: \end{align}
533:
534: Equations~(\ref{eq:ORalpha}--\ref{eq:ORalphbetcc}) can be written
535: in matrix notation as
536: \begin{align}
537: \mathbf{UU}^{\dagger}-\mathbf{VV}^{\dagger}=\mathbf{U}^{*}\mathbf{U}^{T}-
538: \mathbf{V}^{*}\mathbf{V}^{T} & = \mathbf{I}_{N_{a}} \label{eq:UUVV}\,,\\
539: \mathbf{W}\mathbf{W}^{\dagger}-\mathbf{X}\mathbf{X}^{\dagger}=
540: \mathbf{W}^{*}\mathbf{W}^{T}-\mathbf{X}^{*}\mathbf{X}^{T}
541: & = -\mathbf{I}_{N_{b}}\,,\\
542: \mathbf{UW}^{T}-\mathbf{VX}^{T}=\mathbf{WU}^{T}-\mathbf{XV}^{T}
543: & = \mathbf{0}\,,\\
544: \mathbf{W}^{*}\mathbf{U}^{\dagger}-\mathbf{X}^{*}\mathbf{V}^{\dagger}=
545: \mathbf{U}^{*}\mathbf{W}^{\dagger}-\mathbf{V}^{*}\mathbf{X}^{\dagger}
546: & = \mathbf{0}\,,
547: \label{eq:WUXV}
548: \end{align}
549: where $\mathbf{I}_{N_{a}}$ ($\mathbf{I}_{N_{b}}$) is the $N_{a}\times N_{a}$
550: ($N_{b}\times N_{b}$) unit matrix. These relations can be put into
551: a more compact form by defining the matrix
552: \begin{equation}
553: \mathbf{1}_{p}=\begin{pmatrix}
554: \mathbf{I}_{N_{a}} & \mathbf{0}\\
555: \mathbf{0} & -\mathbf{I}_{N_{b}}\end{pmatrix},
556: \label{eq:1p}
557: \end{equation}
558: in terms of which Eqs.~(\ref{eq:UUVV}--\ref{eq:WUXV}) can be rewritten as
559: \begin{equation}
560: \mathbf{T}\mathbf{1}_{p}\mathbf{T}^{\dagger}=\mathbf{1}_{p}\,.
561: \label{eq:T1pT+}
562: \end{equation}
563: Since $\mathbf{1}_{p}\mathbf{1}_{p}=\mathbf{I}_{N_{a}+N_{b}}$, it
564: can be shown, after simple algebraic transformations, that
565: \begin{equation}
566: \mathbf{T}^{\dagger}\mathbf{1}_{p}\mathbf{T}=\mathbf{1}_{p}\,.
567: \label{eq:T+1pT}\end{equation}
568: As a result we have four additional (though not independent) sets
569: of orthogonality equations,
570: \begin{align}
571: \sum_{n=1}^{N_{a}}u_{ni}u_{nj}^{\ast}-\sum_{n=1}^{N_{b}}w_{ni}^{\ast}w_{nj}
572: & = \delta_{ij}\,,\\
573: \sum_{n=1}^{N_{a}}v_{ni}v_{nj}^{\ast}-\sum_{n=1}^{N_{b}}x_{ni}^{\ast}x_{nj}
574: & = -\delta_{ij}\,,\\
575: \sum_{n=1}^{N_{a}}v_{ni}u_{nj}^{\ast}-\sum_{n=1}^{N_{b}}x_{ni}^{\ast}w_{nj}
576: & = 0\,,\\
577: \sum_{n=1}^{N_{a}}u_{ni}v_{nj}^{\ast}-\sum_{n=1}^{N_{b}}w_{ni}^{\ast}x_{nj}
578: & = 0\,.
579: \end{align}
580: Since transformations \eqref{eq:alpha} and \eqref{eq:beta} diagonalize
581: the spin wave Hamiltonian \eqref{eq:Hsw}, the quasi-particles will
582: obey the following commutation relations:
583: \begin{align}
584: \left[\alpha_{n},H_{\text{sw}}\right]
585: & = \omega_{n}^{(\alpha)}\alpha_{n}\,,\label{eq:alphaH}\\
586: \left[\beta_{n},H_{\text{sw}}\right] & = \omega_{n}^{(\beta)}\beta_{n}\,,\\
587: \left[\alpha_{n}^{\dagger},H_{\text{sw}}\right]
588: & = -\omega_{n}^{(\alpha)}\alpha_{n}^{\dagger}\,,\\
589: \left[\beta_{n}^{\dagger},H_{\text{sw}}\right]
590: & = -\omega_{n}^{(\beta)}\beta_{n}^{\dagger}\,.\label{eq:betapH}
591: \end{align}
592: [Notice that we have relabeled the positive eigenvalues introduced
593: in \eqref{eq:diag} to $\omega_{1}^{(\alpha)},\ldots,\omega_{N_{a}}^{(\alpha)},
594: \omega_{1}^{(\beta)},\ldots,\omega_{N_{b}}^{(\beta)}$.]
595: From Eqs.~\eqref{eq:alphaH} and \eqref{eq:alpha} we can define an eigenvalue
596: matrix equation in the usual form, namely,
597: \begin{equation}
598: \begin{pmatrix}
599: \mathbf{K}^{a} & -\mathbf{\Delta}\\
600: \mathbf{\Delta}^{T} & -\mathbf{K}^{b}\end{pmatrix}
601: \begin{pmatrix}
602: \bar{u}_{n}\\
603: \bar{v}_{n}\end{pmatrix}
604: =\omega_{n}^{(\alpha)}
605: \begin{pmatrix}
606: \bar{u}_{n}\\
607: \bar{v}_{n}
608: \end{pmatrix},
609: \label{eq:eiguv}
610: \end{equation}
611: where the column vectors $\bar{u}_{n}$ and $\bar{v}_{n}$ contain
612: the coefficients $u_{ni}$ and $v_{ni}$, respectively. From
613: Eq.~\eqref{eq:betapH} and the complex conjugate of Eq.~\eqref{eq:beta},
614: a similar eigenvalue matrix equation can be defined,
615: \begin{equation} \label{eq:eigwx}
616: \begin{pmatrix}
617: \mathbf{K}^{a} & -\mathbf{\Delta}\\
618: \mathbf{\Delta}^{T} & -\mathbf{K}^{b}
619: \end{pmatrix}
620: \begin{pmatrix}
621: \bar{w}_{n}^{\ast}\\
622: \bar{x}_{n}^{\ast}
623: \end{pmatrix}
624: =-\omega_{n}^{(\beta)}
625: \begin{pmatrix}
626: \bar{w}_{n}^{\ast}\\
627: \bar{x}_{n}^{\ast}
628: \end{pmatrix}.
629: \end{equation}
630: Defining the matrices
631: \begin{equation}
632: \mathbf{\Omega}_{\alpha}=
633: \diag(\omega_{1}^{(\alpha)},\ldots,\omega_{N_{a}}^{(\alpha)})\,,
634: \end{equation}
635: and
636: \begin{equation}
637: \mathbf{\Omega}_{\beta}=
638: \diag(\omega_{1}^{(\beta)},\ldots,\omega_{N_{b}}^{(\beta)})\,,
639: \end{equation}
640: and recalling definition \eqref{eq:T} for $\mathbf{T}$ and definition
641: \eqref{eq:1p} for $\mathbf{1}_{p}$, Eqs.~\eqref{eq:eiguv} and \eqref{eq:eigwx}
642: can be combined into a single equation,
643: \begin{equation}\label{eq:eigprob}
644: \mathbf{D}\mathbf{1}_{p}\mathbf{T}^{\dagger}=\mathbf{T}^{\dagger}
645: \begin{pmatrix}
646: \mathbf{\Omega}_{\alpha} & \mathbf{0}\\
647: \mathbf{0} & -\mathbf{\Omega}_{\beta}
648: \end{pmatrix}.
649: \end{equation}
650: This can be made more compact still by defining the matrix
651: $\mathbf{\Omega}=\diag(\omega_{1}^{(\alpha)},\ldots,
652: \omega_{N_{a}}^{(\alpha)},\omega_{1}^{(\beta)},\ldots,
653: \omega_{N_{b}}^{(\beta)})$, such that
654: \begin{equation}
655: \mathbf{D}\mathbf{1}_{p}\mathbf{T}^{\dagger}=
656: \mathbf{T}^{\dagger}\mathbf{\Omega}\mathbf{1}_{p}\,.
657: \label{DST-TSO}
658: \end{equation}
659: From Eq.~\eqref{DST-TSO} and the relations \eqref{eq:T1pT+} and
660: \eqref{eq:T+1pT}, it is not difficult to show that
661: \begin{equation}
662: \mathbf{1}_{p}\mathbf{T}\mathbf{1}_{p}\mathbf{D}
663: \mathbf{1}_{p}\mathbf{T}^{\dagger}\mathbf{1}_{p}=\mathbf{\Omega}\,.
664: \end{equation}
665: Thus, solving the eigenvalue problem defined by Eq.~\eqref{eq:eigprob}
666: under the constraint \eqref{eq:T1pT+} is equivalent to finding a transformation
667: $\mathbf{T}$ which satisfies Eq.~\eqref{eq:diag}, and where, obviously,
668: \begin{equation}
669: \mathbf{1}_{p}\mathbf{T}^{\dagger}\mathbf{1}_{p}=\mathbf{T}^{-1}\,.
670: \end{equation}
671: According to Eq.~\eqref{eq:rot}, the diagonalized form of the spin wave
672: Hamiltonian is obtained as
673: \begin{equation}
674: \begin{split}
675: H_{\text{sw}} & = (\{ a\},\{ b^{\dagger}\})\mathbf{D}
676: (\{ a\},\{ b^{\dagger}\})^{\dagger}\nonumber \\
677: & = (\{\alpha\},\{\beta^{\dagger}\})\mathbf{\Omega}
678: (\{\alpha\},\{\beta^{\dagger}\})^{\dagger}\nonumber \\
679: & = \sum_{n=1}^{N_{a}}\omega_{n}^{(\alpha)}\alpha_{n}\alpha_{n}^{\dagger}
680: +\sum_{n=1}^{N_{b}}\omega_{n}^{(\beta)}\beta_{n}^{\dagger}\beta_{n}\,.
681: \end{split}
682: \end{equation}
683:
684: The conclusion of the above discussion is that the operator transformation
685: given by Eqs.~\eqref{eq:alpha} and \eqref{eq:beta} diagonalizes
686: the spin wave Hamiltonian \eqref{eq:haf} and that $\alpha^{\dagger}$
687: and $\beta^{\dagger}$ are the quasi-particles (with bosonic character)
688: associated with the low energy excitations of the antiferromagnetic
689: Heisenberg Hamiltonian for quantum spins in a site diluted honeycomb
690: lattice (needless to say the above description applies to other lattices
691: as well).
692:
693: Using Eq.~\eqref{eq:rot} it is possible to write any average of the
694: initial bosonic operators in terms of the quasi-particle operators
695: $\alpha$ and $\beta$. The simplest example is the staggered magnetization
696: $M_{z}^{\text{stagg}}$ at $T=0$ given by
697: \begin{equation} \label{eq:magz}
698: \begin{split}
699: M_{z}^{\text{stagg}} & = \Biggl\langle \sum_{i\in A}S_{i}^{a,z}
700: -\sum_{i\in B}S_{i}^{b,z}\Biggr\rangle \\
701: & = \bigl(N_{a}+N_{b}\bigr)\bigl(S-\delta m_{z}\bigr)\,,
702: \end{split}
703: \end{equation}
704: where
705: \begin{equation} \label{eq:deltamz}
706: \delta m_{z}=\sum_{n=1}^{N_{a}}\delta m_{z}^{(n,\alpha)}
707: +\sum_{n=1}^{N_{b}}\delta m_{z}^{(n,\beta)}\,,
708: \end{equation}
709: with
710: \begin{align}
711: \delta m_{z}^{(n,\alpha)}
712: & = \frac{1}{N_{a}+N_{b}}\sum_{i\in B}|v_{ni}|^{2}\,, \label{eq:deltamzalph}
713: \intertext{and}
714: \delta m_{z}^{(n,\beta)}
715: & = \frac{1}{N_{a}+N_{b}}\sum_{i\in A}|w_{ni}|^{2}.\label{eq:deltamzbet}
716: \end{align}
717:
718:
719: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
720: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
721: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
722: \section{Numerical details}
723: \label{sec:numeric}
724:
725: The formalism developed in the above section is based on the existence
726: of the matrix $\mathbf{T}$ and, naturally, on the possibility of
727: finding it by some numerical procedure. In this work we have used two
728: independent methods to compute the transformation matrix and the associated
729: eigenenergies. Both methods agree with each other within the numerical
730: accuracy of the calculation. One of them is based on a Cholesky decomposition
731: and gives the $\mathbf{T}^{-1}$ matrix directly, whereas the other
732: solves the eigenvalue problem defined by Eq.~\eqref{eq:eigprob},
733: computing the matrix $\mathbf{T}^{\dagger}$ and from this the
734: matrix $\mathbf{T}$.
735:
736: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
737: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
738: \subsection{Cholesky Decomposition method}
739: \label{Cholesky}
740:
741: As shown by Colpa,\cite{colpa} so long as the grand dynamical matrix
742: is positive definite, a simple algorithm exists for determining $\mathbf{T}$.
743: A hermitian (or symmetric) matrix is positive definite if all its
744: eigenvalues are positive. By definition the quasi-particles $\alpha^{\dagger}$
745: and $\beta^{\dagger}$ have positive or zero excitation energy. As
746: will be shown in Subsect.~\ref{sub:Zm}, the zero energy excitations
747: are associated with spin rotations, which cost zero energy due to
748: the spin rotational symmetry of the isotropic Heisenberg model. So,
749: provided that $h_{a}\geq1^{+}$, all eigenvalues are positive and the
750: grand dynamical matrix is positive definite. The algorithm is implemented
751: in three major steps:
752: \begin{enumerate}
753: \item for $\mathbf{D}$ positive definite a Cholesky decomposition can be
754: performed \cite{recipes} and we have
755: $\mathbf{D}=\mathbf{Q}\mathbf{Q}^{\dagger}$,
756: where $\mathbf{Q}$ is an upper triangular matrix. The existence of
757: a Cholesky decomposition guarantees that the problem is positive definite;
758: \item it can be proved that there exists a unitary transformation $\mathbf{Y}$
759: such that $\mathbf{Y}^{\dagger}(\mathbf{Q}\mathbf{1}_{p}^{(ab)}
760: \mathbf{Q}^{\dagger})\mathbf{Y}=\mathbf{1}_{p}^{(ab)}\diag(\omega_{1},
761: \ldots,\omega_{N_{a}},\omega_{N_{a}+1},\ldots\omega_{N_{a}+N_{b}})$;
762: \item finally, it can be proved that $\mathbf{T}^{-1}=\mathbf{Q}^{-1}
763: \mathbf{Y}\diag(\sqrt{\omega_{1}},\ldots,\sqrt{\omega_{N_{a}+N_{b}}})$.
764: \end{enumerate}
765:
766: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
767: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
768: \subsection{Bogoliubov-Valatin Transformation method}
769: \label{sub:BV}
770:
771: The nonhermitian eigenproblem defined by Eq.~\eqref{eq:eigprob}
772: can be solved with standard numerical algorithms. Here we have used
773: subroutines of the LAPACK library. It should be noted that the resultant
774: eigenvectors do not provide the required $\mathbf{T}^{\dagger}$ matrix
775: directly. After diagonalization the eigenvectors have to be normalized
776: such that they satisfy Eqs.~\eqref{eq:ORalpha} and \eqref{eq:ORbeta}
777: for $n=m$. Degenerate eigenvectors%
778: \footnote{Although degeneracy is removed by disorder, we have to handle it if
779: we want our algorithm to be valid in the undiluted case.%
780: } have to be carefully analyzed because the LAPACK subroutines we have
781: used do not guarantee that they satisfy Eq.~\eqref{eq:T1pT+}
782: (though Eqs.~\eqref{eq:ORalpha} to \eqref{eq:ORalphbetcc} are satisfied
783: by default for $n\neq m$). The algorithm is implemented as follows:
784: \begin{enumerate}
785: \item the matrix $\mathbf{D}\mathbf{1}_{p}^{(ab)}$ is reduced to an upper
786: Hessenberg form $\mathbf{H}$ by an orthogonal transformation $\mathbf{Y}$,
787: \emph{i.e.}, $\mathbf{H}=\mathbf{YD1}_{p}^{(ab)}\mathbf{Y}^{\dagger}$ (LAPACK
788: subroutines DGEHRD and DORGHR);
789: \item the eigenvalues of the upper Hessenberg matrix (the same as those
790: of $\mathbf{D}\mathbf{1}_{p}^{(ab)}$) and the matrices $\mathbf{Q}$
791: and $\mathbf{Z}$ from the Schur decomposition $\mathbf{H}=
792: \mathbf{ZQZ}^{\dagger}$, where $\mathbf{Q}$ is an upper quasi-triangular
793: matrix (the Schur form), and $\mathbf{Z}$ is the orthogonal matrix of Schur
794: vectors, are computed (LAPACK subroutine DHSEQR);
795: \item the right eigenvectors of the upper quasi-triangular matrix $\mathbf{Q}$
796: are computed and multiplied by $\mathbf{Y}^{\dagger}\mathbf{Z}$,%
797: \footnote{Subroutine DHSEQR already gives the product
798: $\mathbf{Y}^{\dagger}\mathbf{Z}$ instead of $\mathbf{Z}$.%
799: } giving the right eigenvectors of $\mathbf{D}\mathbf{1}_{p}^{(ab)}$,
800: whose matrix form we name $\mathbf{\widetilde{T}}^{\dagger}$ (LAPACK
801: subroutines DTREVC);
802: \item degenerate column eigenvectors of $\mathbf{\widetilde{T}}^{\dagger}$
803: are arranged in linear combinations such that they satisfy
804: Eq.~\eqref{eq:T1pT+};
805: \item nondegenerate column eigenvectors of $\mathbf{\widetilde{T}}^{\dagger}$
806: are normalized so as to satisfy the orthogonality relations of Eqs.
807: \eqref{eq:ORalpha} and \eqref{eq:ORbeta}, giving matrix
808: $\mathbf{T}^{\dagger}$;
809: \item the positive eigenvalues and respective eigenvectors are identified
810: with $\alpha$ modes, and the negative ones with the $\beta$ modes;
811: \item eigenvalues and eigenvectors are sorted such that matrix
812: $\mathbf{T}^{\dagger}$ has the form defined in Eq.~\eqref{eq:T}.
813: \end{enumerate}
814:
815: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
816: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
817: \subsection{Zero modes}
818: \label{sub:Zm}
819:
820: It is well known that the clean and isotropic limit of Hamiltonian
821: \eqref{eq:haf} has two zero-energy excitations (Goldstone bosons),
822: whose momenta can be determined from Eq.~\eqref{eq:Hkdiag}. These
823: gapless modes are a consequence of the fact that the ground state
824: spontaneously breaks the rotational symmetry of the Hamiltonian in
825: spin space. It can be shown\cite{ConcSolid63} that these zero-energy
826: modes have divergent amplitudes. In two and three dimensions the quantum
827: corrections to the staggered magnetization (at zero temperature) are
828: finite, meaning that the divergence associated with the zero energy
829: modes is integrable. We note, however, that if the mean square amplitudes
830: of the differences between the two $x$- and $y$-components, given by
831: \begin{align}
832: \frac{1}{N_{a}+N_{b}}&\left(\sum_{i\in A}S_{i}^{a,x}
833: -\sum_{i\in B}S_{i}^{b,x}\right)\label{eq:magx}
834: \intertext{and}
835: \frac{1}{N_{a}+N_{b}}&\left(\sum_{i\in A}S_{i}^{a,y}
836: -\sum_{i\in B}S_{i}^{b,y}\right)\,,\label{eq:magy}
837: \end{align}
838: are computed, we immediately find divergent behavior.\cite{AND52}
839: The same divergent behavior is also found if we try to get the staggered
840: magnetization \eqref{eq:magz} from a finite-size scaling procedure,
841: doing summations in $\mathbf{k}\text{-space}$ for finite-size systems,
842: including all momenta of the Brillouin zone.
843:
844: As shown by Anderson,\cite{AND52,ConcSolid63} this does not mean that
845: the spin wave approximation is breaking down and that the system
846: has no LRO. What it means is that these divergences are related to
847: the zero point motion of the Goldstone modes, and their presence is
848: required to exist since in a finite-size system one cannot have solutions
849: that break the spin rotational symmetry of the problem.\cite{Hulthen1938}
850: The presence of a \emph{broken symmetry ground state} is made possible
851: if we analyze the $H_{\mathbf{0}}$ term in Eq.~\eqref{eq:Hk}, from
852: which the Goldstone modes arise. This term cannot be diagonalized
853: through any Bogoliubov-Valatin transformation. Actually, it has a
854: continuum spectrum starting from the zero energy ground state (see
855: Appendix \ref{sec:App-diagHk0}). Using this continuum of states we
856: can form a wave packet centered around some prefixed orientation in
857: spin space, with the property of having both a finite staggered magnetization,
858: and a mean square roots of the quantities \eqref{eq:magx} and \eqref{eq:magy}
859: scaling with $1/N^{\frac{1}{2}-\alpha}$, with $\alpha>0$, as long
860: as we pay some extra energy. In addition, it can be shown (see Appendix
861: \ref{sec:App-Anderson}) that this extra energy scales as
862: $1/N^{\frac{1}{2}+\alpha}$, being negligible in the thermodynamic limit.
863: Thus it is a suitable approximation to form the above mentioned wave packet
864: from the solutions of $H_{\mathbf{0}}$, and to study the energy and the
865: zero point motion of all other normal modes within a time interval smaller
866: than that needed for the zero-energy wave packet to disrupt the coherence of
867: the unidirectional state.\cite{ConcSolid63} The understanding of
868: the role played by Goldstone bosons in finite-size systems is crucial
869: for computing well defined quantities in the thermodynamic limit from
870: calculations in finite-size lattices. As a practical example, the
871: staggered magnetization can be obtained from the finite-size calculations
872: if the Goldstone zero point motion contributions are subtracted, because
873: the $H_{\mathbf{0}}$ solutions were already used to form the starting
874: broken symmetry state. From this procedure we get exactly the same
875: value as from the convergent integral in the continuum limit.
876:
877: The above discussion now needs to be carried on to the diluted case,
878: where the above aspects are more delicate than in the nondisordered
879: case. In the presence of dilution, it is easy to verify that there
880: is at least one zero mode in Eq.~\eqref{eq:eiguv} in the isotropic
881: case. This nontrivial solution with zero energy satisfies the equation
882: \begin{equation} \label{eq:zm}
883: \begin{pmatrix}
884: \mathbf{K}^{a} & -\mathbf{\Delta}\\
885: \mathbf{\Delta}^{T} & -\mathbf{K}^{b}
886: \end{pmatrix}
887: \begin{pmatrix}
888: \bar{c}\\
889: \bar{c}\end{pmatrix} =0,
890: \end{equation}
891: with all the amplitudes constant. To prove that this is indeed an
892: eigenstate we only need to remember definitions
893: (\ref{eq:Ka}--\ref{eq:DeltaT}) of matrices $\mathbf{K}^{a}$, $\mathbf{K}^{b}$
894: and $\mathbf{\Delta}$, and check that the following equalities
895: always hold:
896: \begin{align}
897: K_{ii}^{a} & = \!\sum_{j=N_{a}}^{N_{a}+N_{b}}\!\Delta_{ij}\,,\\
898: K_{ii}^{b} & = \sum_{j=1}^{N_{a}}\Delta_{ij}^{T}\,.
899: \end{align}
900:
901: In terms of quasi-particle excitations, the eigenvector defined by
902: Eq.~\eqref{eq:zm} can be expressed as%
903: \footnote{The negligence of normalization doesn't change the conclusions we
904: will arrive. Actually, this modes will be identified with the Goldstone
905: modes of the diluted system, which, as in the clean limit, can have
906: divergent amplitude.}
907: \begin{align}
908: \alpha_{0}^{\dagger}&\propto\sum_{i=1}^{N_{a}}a_{i}^{\dagger}
909: +\sum_{i=1}^{N_{b}}b_{i}\,,\label{eq:zm-alph-ab}
910: \intertext{in the case of an $\alpha$-type excitation, and}
911: \beta_{0}^{\dagger}&\propto\sum_{i=1}^{N_{a}}a_{i}
912: +\sum_{i=1}^{N_{b}}b_{i}^{\dagger} \label{eq:zm-beta-ab}\,,
913: \end{align}
914: if it is a $\beta$-type excitation. Recalling the approximate
915: expressions in Eq.~\eqref{eq:Sapprox} for the operators $S_{i}^{(a,b)+}$
916: and $S_{i}^{(a,b)-}$ in terms of bosonic operators $a$ and $b$,
917: Eqs.~\eqref{eq:zm-alph-ab} and \eqref{eq:zm-beta-ab} can be rewritten as
918: \begin{align}
919: \alpha_{0}^{\dagger} & \propto S_{\text{tot}}^{-}\,, \label{eq:zm-alph-Stot-}\\
920: \beta_{0}^{\dagger} & \propto S_{\text{tot}}^{+}\,. \label{eq:zm-beta-Stot+}
921: \end{align}
922: Thus, excitations $\alpha_{0}^{\dagger}$ and $\beta_{0}^{\dagger}$
923: are precisely the Goldstone bosons associated with the broken
924: continuous symmetry of spin rotation in the diluted system.
925:
926: As will be shown in Subsect.~\ref{sub:Finite-size-scaling}, the thermodynamic
927: limit of the staggered magnetization for the diluted system will be
928: obtained from a finite-size scaling analysis. As we have started from
929: a broken symmetry ground state (the wave packet), which is a direct
930: consequence of Eqs.~\eqref{eq:NsGA} and \eqref{eq:NsGB}, we would
931: proceed as in the clean limit and neglect the contributions of $\alpha_{0}$
932: and $\beta_{0}$ modes. However, although in the undiluted case
933: the number of Goldstone modes is always two, when dilution is present
934: this number can either be one or two, in a finite size lattice. The
935: reason why this is so is that operators $S_{\text{tot}}^{-}$ and
936: $S_{\text{tot}}^{+}$ do not always represent independent excitations,
937: \emph{i.e.}, they do not always commute. Naturally $S_{\text{tot}}^{-}$
938: and $S_{\text{tot}}^{+}$ never commute strictly speaking because
939: \begin{equation}
940: [S_{\text{tot}}^{+},S_{\text{tot}}^{-}]=2S_{\text{tot}}^{z}\,.
941: \label{eq:commutS+S-}
942: \end{equation}
943: Nevertheless, in the clean limit we can easily convince ourselves
944: that the expectation value of $S_{\text{tot}}^{z}$ is always zero,
945: and, as $S_{\text{tot}}^{z}$ is a constant of the motion, commutator
946: \eqref{eq:commutS+S-} will always be zero. To get the value of the
947: commutator \eqref{eq:commutS+S-} in the presence of dilution we make
948: use of Eq.~\eqref{eq:Sapprox}, from which one finds
949: \begin{equation}
950: [S_{\text{tot}}^{+},S_{\text{tot}}^{-}]\propto N_{a}-N_{b}\,.
951: \label{eq:commutS+S-dilute}
952: \end{equation}
953: Now it is easily seen that one can have one or two Goldstone modes
954: in a finite-size diluted system: if the number of undiluted sites
955: in each sublattice is the same ($N_{a}=N_{b}$) there will be two
956: Goldstone modes; otherwise, if $N_{a}\neq N_{b}$, there will be only
957: one. Applying to this case the reasoning used for the undiluted
958: case, we should then neglect the contributions of the existent Goldstone
959: modes.
960:
961: As the system size increases, the fluctuations relative to the zero
962: mean value of $N_{a}-N_{b}$ should scale as $1/\sqrt{N_{a}+N_{b}}$,
963: statistically speaking. Therefore the difference $N_{a}-N_{b}$ is
964: again zero in the thermodynamic limit and the system has two zero
965: energy excitations. This situtation cannot be achieved in finite
966: size lattices, unless we restrict ourselvess to cases where the disordered
967: realizations are constrained to obey the condition $N_{a}=N_{b}$, being
968: clear that the staggered magnetization in the thermodynamic limit
969: cannot depend on this restriction. We stress, however, that without
970: this restriction the conclusions drawn from finite-size lattices would
971: be different if we had accepted all sorts of disordered lattice realizations.
972: This difference is due to the contribution of the {}``quasi-divergent''
973: energy modes that emerge when $N_{a}\neq N_{b}$. We will get back
974: to this point in Sect.~\ref{sec:results}, presenting numerical evidence
975: for what we have just analysed.
976:
977: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
978: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
979: \subsection{Cluster formation and periodic boundary conditions}
980:
981: The study of diluted lattices requires the concept of \emph{largest
982: cluster}, and therefore some care is required in constructing the
983: effective lattice where the quantum problem is to be solved. Since
984: we are interested in dilution, the algorithms discussed in Subsecs.
985: \ref{Cholesky} and \ref{sub:BV} are to be implemented not on all
986: occupied lattice sites, but only on the sites defined by the largest
987: connected cluster of spins, since in the thermodynamic limit a finite
988: magnetization cannot exist if one is below the percolation critical
989: threshold $p_{\text{c}}$. The dilution is induced in the lattice by diluting
990: any site with probability $1-p$. For $p=1$ there is no dilution
991: at all. When $p=p_{\text{c}}$ a classical percolation transition occurs
992: in the thermodynamic limit preventing the existence of magnetic long
993: range order in the system. According to Suding and Ziff,\cite{SZ99}
994: $p_{\text{c}}=0.697043(3)$ in the honeycomb lattice. Here we use
995: $p_{\text{c}}=0.697043$.
996:
997: We work with finite size lattices where periodic boundary conditions
998: (p.b.c.) are implemented as defined in Fig.~\ref{cap:honey}. In
999: Fig.~\ref{cap:honey} the links on the border are labeled according to which
1000: site they connect to. The lattices are characterized
1001: by their linear dimension $L$ ($L=3$ in Fig.~\ref{cap:honey}). The
1002: total number of sites for a given $L$ is $2L^{2}$.%
1003: \begin{figure}
1004: \begin{center}\includegraphics[
1005: width=8cm]{fig1-hsd.eps}\end{center}
1006:
1007: \caption{(color online) A finite size honeycomb lattice showing the periodic
1008: boundary conditions used in the numerical calculations.}
1009:
1010: \label{cap:honey}
1011: \end{figure}
1012:
1013: The algorithm starts by identifying the largest cluster, for rigid
1014: boundary conditions (this is, with no p.b.c.). As in Ref.~\onlinecite{MNC04},
1015: it is only after the largest cluster is found that we apply p.b.c. to
1016: the original lattice, checking whether there are new sites belonging to
1017: the largest cluster. As previously discussed in Subsect.~\ref{sub:Zm},
1018: only clusters with $N_{a}=N_{b}$ are to be used, so we reject all disordered
1019: lattice realizations in which $N_{a}\neq N_{b}$.%
1020: \footnote{This procedure is a highly inefficient one,
1021: because only a small fraction ($\sim6$\% for $L=14$) of disordered lattice
1022: realization are accepted. Nevertheless, the amount of time spent finding
1023: clusters with $N_{a}=N_{b}$ is a small percentage ($\sim15$\% and $\sim6$\%
1024: for $L=14$ in the \emph{Cholesky decomposition method} and
1025: \emph{Bogoliubov-Valatin transformation method}, respectively) of the time
1026: consumed by the diagonalization subroutines.%
1027: } Finally, the eigenvalue problem is solved for the final cluster using
1028: the aforementioned algorithms. In Fig.~\ref{cap:perc} we show an
1029: example of a disorder realization and the corresponding cluster labeling
1030: process at $p=p_{\text{c}}$. The larger cluster found for rigid boundary
1031: conditions can be seen in panel (\textbf{c}) of Fig.~\ref{cap:perc}.
1032: After p.b.c. implementation the final cluster has a larger number
1033: of elements (panel (\textbf{d}) in Fig.~\ref{cap:perc}).%
1034: \begin{figure}
1035: \begin{center}\includegraphics[%
1036: scale=0.36]{fig2-hsd.eps}\end{center}
1037:
1038: \caption{(color online) An example, for a disorder realization in a lattice
1039: with $L=10$, of the cluster formation. In \textbf{(a)} is the original
1040: lattice; in \textbf{(b)} each site is chosen to be diluted with probability
1041: $1-p_{\text{c}}$ with $p_{\text{c}}=0.697043$; in \textbf{(c)} the several (12
1042: in this case) clusters with rigid boundary conditions have been determined;
1043: in \textbf{(d)} the larger cluster found in \textbf{(c)} is augmented
1044: by the periodic boundary conditions.}
1045:
1046: \label{cap:perc}
1047: \end{figure}
1048:
1049: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1050: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1051: \subsection{Finite-size scaling}
1052: \label{sub:Finite-size-scaling}
1053:
1054: The eigenvalue problem determines all the eigenvalues and eigenfunctions
1055: for the cluster, and from these the corrections to the staggered magnetization
1056: are computed according to Eq.~\eqref{eq:deltamz}. For a given $p$
1057: value, $N_{\text{rz}}$ disordered lattice realizations with $N_{a}=N_{b}$
1058: are performed, leading to an average staggered magnetization density
1059: $m_{\text{av}}$
1060: \begin{equation}
1061: \label{eq:mav}
1062: m_{\text{av}}(p,L) = \frac{1}{N_{\text{rz}}}
1063: \sum_{i=1}^{N_{\text{rz}}}\frac{M_{z}^{\text{stagg,}i}}{N_{m}^{i}}\,,
1064: \end{equation}
1065: where $M_{z}^{\text{stagg,}i}$ is the value of Eq.~\eqref{eq:magz},
1066: and $N_{m}^{i}$ is the total number of magnetic (undiluted) sites
1067: in the lattice, for the given disorder realization $i$. Although
1068: $m_{\text{av}}$ does not depend explicitly on $L$, the sizes of the clusters
1069: are determined by $L$, and therefore different $L$'s lead to different
1070: values for Eq.~\eqref{eq:mav}. With this definition we will be able to
1071: identify $m_{\text{av}}(p,L\rightarrow\infty)$ with the ordered magnetic
1072: moment magnitude per magnetic ion measured in neutron diffraction
1073: experiments.
1074:
1075: From Eq.~\eqref{eq:magz} it is easily seen that $m_{\text{av}}$ can be
1076: expressed as the average product of two different contributions, one
1077: purely classical ($m_{\text{cl}}^{i}$) and the other purely quantum
1078: ($m_{\text{qm}}^{i}$),
1079: \begin{equation}
1080: m_{\text{av}}(p,L)=\frac{1}{N_{\text{rz}}}\sum_{i=1}^{N_{\text{rz}}}
1081: m_{\text{cl}}^{i}m_{\text{qm}}^{i}\,,
1082: \label{eq:mavMclMqm}
1083: \end{equation}
1084: where we used the notation $m_{\text{cl}}^{i}=
1085: \frac{N_{\text{c}}^{i}}{N_{m}^{i}}$ for the classical factor,
1086: with $N_{\text{c}}^{i}=N_{a}^{i}+N_{b}^{i}$, and
1087: $m_{\text{qm}}^{i}=S-\delta m_{z}^{i}$ for the quantum mechanical factor.
1088: The quantum contribution is simply the staggered magnetization density
1089: of the larger cluster found in the disorder realization $i$. It would
1090: be $S$ in the N\'{e}el state but it is reduced by $\delta m_{z}^{i}(p,L)$
1091: due to quantum fluctuations, whose strength depend on dilution $p$
1092: and lattice size $L$. If LRO is present we can assume that the sublattice
1093: magnetization, or equivalently the staggered magnetization, is a self-averaging
1094: quantity, as was shown to happen in the square lattice case.\cite{San02}
1095: Thus, in the thermodynamic limit of $m_{\text{av}}$ each disorder realization
1096: $m_{\text{qm}}^{i}$ can then be replaced by its infinite-size-extrapolated
1097: average, which we denote by $m_{\text{qm}}$,
1098: \begin{equation}
1099: m_{\text{av}}(p,L\rightarrow\infty)=m_{\text{cl}}m_{\text{qm}}\,.
1100: \label{eq:mavLinfty}
1101: \end{equation}
1102: The classical factor now assumes the standard form for the order
1103: parameter of the classical percolation problem,
1104: \begin{equation}
1105: m_{\text{cl}}=
1106: \left\langle \frac{N_{\text{c}}}{N_{m}}\right\rangle _{L\rightarrow\infty}\,,
1107: \label{eq:mcl}
1108: \end{equation}
1109: which is zero for $p\leq p_{\text{c}}$. Therefore a quantum critical point
1110: can only exist above $p_{\text{c}}$ if $m_{\text{qm}}=0$ for some
1111: $p^{*}>p_{\text{c}}$.
1112: To find $m_{\text{qm}}$ we need to compute the average infinite-size value
1113: of the quantum corrections $\delta m_{z}^{\infty}$ from our finite
1114: size calculations. We show that finite-size scaling can be found for
1115: this quantity, from which results holding in the thermodynamic limit
1116: can be obtained. In our study the size of the largest connected cluster
1117: $N_{a}+N_{b}$ is not fixed, instead the linear dimension of the lattice
1118: $L$ is. As shown for the square lattice,\cite{San02} the alternative
1119: approach where the percolating cluster size is fixed leads to the
1120: same magnetization value in the thermodynamic limit. The finite-size
1121: scaling properties of the quantum correction to the magnetization are
1122: strictly not known for a disordered system at the percolation point.
1123: However, in practice a direct generalization of the pure-system
1124: scaling, using the fractal (Hausdorff) dimensionality, has been
1125: shown to work well.\cite{San02} Hence we will assume
1126: \begin{equation} \label{eq:scaling}
1127: \left\langle \delta m_{z}(p,L)\right\rangle _{N_{\text{rz}}}=
1128: \delta m_{z}^{\infty}+aL^{-D/2}+bL^{-D}\,,
1129: \end{equation}
1130: where $\delta m_{z}^{\infty}$ is the average quantum correction
1131: to the staggered magnetization density in the thermodynamic limit,
1132: and $D$ is the fractal dimension of the cluster, which should have
1133: the universal value $D=91/48$ at $p_{\text{c}}$ (in two dimensions),
1134: as is confirmed for the square and triangular lattices.\cite{PS90}
1135:
1136: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1137: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1138: \subsection{Density of states}
1139: \label{sub:DOSrm}
1140:
1141: The real space diagonalization procedure, either \emph{Bogoliubov-Valatin}
1142: or \emph{Cholesky decomposition}, is very time consuming, preventing
1143: us from accessing large clusters (in the honeycomb lattice $L=16$
1144: is our upper limit). Although for the staggered magnetization density
1145: a finite-size scaling analysis can be done, we cannot easily guess
1146: the thermodynamic limit behaviour of the DOS from results of systems
1147: as small as $L=16$.
1148:
1149: In this work the well known recursion method is used to compute the
1150: average DOS. With this method we can handle lattices as large as $L=128$,
1151: with the advantage that the obtained DOS is not the typical finite
1152: size DOS of a system with $L=128$, but instead a very good approximation
1153: for its thermodynamic limit value, guessed from this finite-size system.
1154: We refer the reader to the paper of R. Haydock\cite{Hayd80} for details
1155: in the case of non-interacting fermionic systems. Being a real space
1156: method the effect of disorder can be easily incorporated. Here we
1157: adopt the formulation introduced in Ref.~\onlinecite{HHT95} for disordered
1158: electronic systems. Further details on the recursion method in relation
1159: to disordered bosonic bilinear systems [such as model Hamiltonian
1160: \eqref{eq:Hsw}] will be presented elsewhere.\cite{Castro05}
1161:
1162: It is worth mentioning that the recursion method has proved to be
1163: a powerful technique even in the presence of interactions.\cite{Hay00}
1164: Actually, the continued fraction representation of the Fourier components
1165: of the one particle propagator, the basis of the recursion method,
1166: is also an essential point in the Pad\'{e} analytical continuation which
1167: usually arises in the many-body problem.\cite{BGM00}
1168:
1169: We define the following set of zero temperature retarded Green's functions
1170: in the standard way,
1171: \begin{equation} \label{eq:GRt}
1172: \begin{split}
1173: G_{ij}^{ab}(t) & = -i\left\langle 0\right|\{ a_{i}^{\dagger}(t),
1174: b_{j}^{\dagger}(0)\}\left|0\right\rangle \Theta(t),\\
1175: G_{ij}^{ba}(t) & = -i\left\langle 0\right|\{ b_{i}(t),a_{j}(0)\}
1176: \left|0\right\rangle \Theta(t),\\
1177: G_{ij}^{aa}(t) & = -i\left\langle 0\right|\{ a_{i}^{\dagger}(t),
1178: a_{j}(0)\}\left|0\right\rangle \Theta(t),\\
1179: G_{ij}^{bb}(t) & = -i\left\langle 0\right|\{ b_{i}(t),b_{j}^{\dagger}(0)\}
1180: \left|0\right\rangle \Theta(t),
1181: \end{split}
1182: \end{equation}
1183: where the notation $\left|0\right\rangle $ is used for the ground
1184: state of the spin wave Hamiltonian \eqref{eq:Hsw}. The Fourier components
1185: of each of the Green's functions in Eq.~\eqref{eq:GRt},
1186: \begin{equation}
1187: G_{ij}(E+i0^{+}) =
1188: \int_{-\infty}^{\infty}dt\, e^{i(E+i0^{+})t}G_{ij}(t),
1189: \label{eq:GRe}
1190: \end{equation}
1191: are the quantities of interest when determining the DOS. Defining
1192: the DOS as
1193: \begin{equation}
1194: \rho(E)=\frac{1}{N_{\text{c}}}\Biggl[\sum_{n=1}^{N_{a}}
1195: \delta(E-\omega_{n}^{(\alpha)})
1196: +\sum_{n=1}^{N_{b}}\delta(E-\omega_{n}^{(\beta)})\Biggr],
1197: \label{eq:DOS}
1198: \end{equation}
1199: it can be easily shown that $\rho(E)$ is given in terms of the Fourier
1200: components of the Green's functions \eqref{eq:GRe} as
1201: \begin{multline} \label{eq:DOSG}
1202: \rho(E) = -\frac{1}{N_{\text{c}}}\frac{1}{\pi}
1203: \text{Im}\Biggl[\sum_{i\in A}G_{ii}^{aa}(E+i0^{+})\\
1204: - \sum_{i\in B}G_{ii}^{bb}(E+i0^{+}) - \sum_{i\in A}G_{ii}^{aa}(-E+i0^{+})\\
1205: +\sum_{i\in B}G_{ii}^{bb}(-E+i0^{+})\Biggr].
1206: \end{multline}
1207: The recursion method gives $\text{Im}[G_{ij}(E+i0^{+})]$ directly,
1208: the imaginary part of the Fourier components defined in
1209: Eq.~\eqref{eq:GRe},\cite{Castro05} from which the DOS is straightforwardly
1210: computed.
1211:
1212:
1213: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1214: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1215: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1216: \section{Results}
1217: \label{sec:results}
1218:
1219: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1220: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1221: \subsection{Larger cluster statistics}
1222: \label{subsec:clusterstat}
1223:
1224: The number of sites in a regular planar lattice goes as the square
1225: of its linear size. In the thermodynamic limit, the same scaling applies
1226: to the largest cluster of the corresponding randomly-site-diluted
1227: lattice. This behaviour persists up to the percolation threshold,
1228: at which point the lattice is dominated by a spanning cluster of \emph{fractal}
1229: dimension. Beyond percolation, individual clusters are no longer extensive:
1230: they each constitute a vanishing fraction of the total number of sites.
1231:
1232: For a honeycomb lattice of size $L$ and dilution level
1233: $x=(1-p)/(1-p_{\text{c}})$, let $P(N_{\text{c}}|L,x)$ denote the probability
1234: that the largest cluster has $N_{\text{c}}$ sites. The average size of
1235: the largest cluster is simply the corresponding first moment:
1236: \begin{equation}
1237: \bar{N}_{\text{c}}(L,x)=\sum_{N_{\text{c}}=1}^{2L^{2}}N_{\text{c}}
1238: P(N_{\text{c}}|L,x).
1239: \label{EQ:Nclave}
1240: \end{equation}
1241: Example probability distributions for the honeycomb lattice are given
1242: in Fig.~\ref{cap:FIG:clustersize}. For small $x$, the distributions
1243: are sharply peaked. As $x\rightarrow1$, they become progressively
1244: broader and develop long tails skewed toward small values of $N_{\text{c}}$
1245: (marking the evolution to a different universal scaling function at
1246: percolation).%
1247: \begin{figure}
1248: \begin{center}\includegraphics[%
1249: scale=0.85]{fig3-hsd.eps}\end{center}
1250:
1251: \caption{\label{cap:FIG:clustersize} (color online) The solid (blue) lines
1252: show the distribution of the number of sites in the largest cluster
1253: of a randomly site-diluted honeycomb lattice. From top to bottom,
1254: the three panels correspond to dilution levels $x=0.1,0.5,0.8$. From
1255: left to right, the peaks correspond to linear system sizes $L=4,5,\ldots,18$.
1256: The (red) vertical dotted lines indicated the average cluster sizes
1257: $\bar{N}_{\text{c}}$, computed as per Eq.~\eqref{EQ:Nclave}.}
1258: \end{figure}
1259:
1260: An effective scaling dimension $D_{\text{eff}}(L,x)$ can be defined
1261: by the relation $\bar{N}_{\text{c}}\sim L^{D_{\text{eff}}}$. Its
1262: evolution with $L$ is plotted in Fig.~\ref{cap:FIG:clusterscaling}.
1263: Note that $D_{\text{eff}}(L,x)$ has two points of attraction in the
1264: limit $L\rightarrow\infty$: $D_{\text{eff}}(L,x<1)\rightarrow2$
1265: and $D_{\text{eff}}(L,1)\rightarrow91/48$. Plotted in the appropriate
1266: reduced coordinates---\emph{viz}., $L^{D}P(N_{\text{c}})$ versus
1267: $L^{-D}N_{\text{c}}$ where $D=2$ below percolation and $D=91/48$
1268: at percolation---the probability distribution tends to either a simple
1269: delta function or the nontrivial curve shown in the inset of
1270: Fig.~\ref{cap:FIG:clusterscaling}.%
1271: \begin{figure}
1272: \begin{center}\includegraphics[%
1273: scale=0.85]{fig4-hsd.eps}\end{center}
1274:
1275: \caption{\label{cap:FIG:clusterscaling} (color online) The effective scaling
1276: dimension of the largest cluster takes one of two values in the
1277: $L\rightarrow\infty$ limit: $D_{\text{eff}}=2$ ($0\leq x<1$) or
1278: $D_{\text{eff}}=91/48$ ($x=1$). For $x\lesssim0.5$, $D_{\text{eff}}$ is close
1279: to its asymptotic value at all systems sizes. When $x$ is close to 1,
1280: very large system sizes are necessary to reach the asymptotic regime.
1281: The figure inset shows the largest-cluster size distribution at percolation
1282: plotted in reduced coordinates. Each curve is computed as a histogram over
1283: $10^{5}$ disorder realizations for system sizes $L=5,6,\ldots,48$.
1284: As $L\rightarrow\infty$, the finite-size results converge to a smooth
1285: scaling function (one not dissimilar from that of the square-lattice
1286: case; see Fig.~2 of Ref.~\onlinecite{San02}).}
1287: \end{figure}
1288:
1289: As can be seen in Fig.~\ref{cap:FIG:clusterscaling} (inset), a long
1290: tail is present for smaller cluster sizes. This enhancement of the
1291: larger cluster size distribution can be understood as a consequence
1292: of the many possible disorder configurations for the same dilution.
1293: That is, we can have various smaller clusters instead of one large
1294: dominant cluster for the same number of diluted sites, though, of
1295: course, these disorder configurations are not so favorable.
1296:
1297: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1298: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1299: \subsection{Finite size scaling analysis}
1300:
1301: We have performed numerical real space diagonalization of model Hamiltonian
1302: \eqref{eq:Hsw}, as described in Sect.~\ref{sec:numeric}, for the
1303: honeycomb and the square lattices. Lattices with sizes $L=5,6,7,8,9,10,11,12,
1304: 13,14,15,16$ (honeycomb) and $L=6,8,10,12,14,16,18,20,22,24,26,28$ (square)
1305: were generated. Averages were taken over $N_{\text{rz}}=10^{5}$ disorder
1306: realizations.%
1307: \footnote{As the simulated lattices have $N=2\times L\times L$ sites, and the
1308: dilution is achieved generating a random number $r\in[0,1]$ at each
1309: lattice site, we need to be careful with the period of the random
1310: number generator when averaging over $10^{5}$ disorder realizations.
1311: In this work we have used the maximally equidistributed combined Tausworthe
1312: generator,\cite{LEc96} as implemented in the \emph{GNU Scientific
1313: Library}. The period of this generator is $2^{88}$ ($\sim10^{26}$). %
1314: }
1315:
1316: In Figure \ref{cap:fssmz} we show, for the honeycomb lattice, the
1317: average quantum correction to the staggered magnetization $\left\langle
1318: \delta m_{z}(p,L)\right\rangle _{N_{\text{rz}}}$,
1319: for various values of dilution $x=(1-p)/(1-p_{\text{c}})$, as a function
1320: of lattice size $L^{-D/2}$. The error bars are much smaller than
1321: the symbols used. The lines are fits to the points using the finite-size
1322: scaling hypotheses \eqref{eq:scaling}. The extrapolated zero abscissa
1323: value gives the average quantum correction to the staggered magnetization
1324: density in the thermodynamic limit $\delta m_{z}^{\infty}(p)$. In
1325: the undiluted case there is an excellent agreement between the real
1326: space diagonalization results (left-triangles) and the reciprocal
1327: space sum (black squares), obtained from the first $\mathbf{k}$-summation
1328: in Eq.~(C9) of Ref.~\onlinecite{PAB04}, thus providing a reliability
1329: test to our algorithms.%
1330: \begin{figure}
1331: \begin{center}\includegraphics[%
1332: scale=0.75]{fig5-hsd.eps}\end{center}
1333:
1334: \caption{\label{cap:fssmz}(color online) Finite size scaling of
1335: $\left\langle \delta m_{z}\right\rangle $ for different values of
1336: $x=(1-p)/(1-p_{\text{c}})$, obtained after $10^{5}$ disorder realizations
1337: of lattices with equal number of sites in each sublattice. Also shown for
1338: $x=1$ is the result obtained when the realized lattices are not constrained
1339: to have $N_{a}=N_{b}$: ({*}) zero modes were subtracted and the highest
1340: amplitude (nonzero) mode (see text) was subtracted if $N_{a}\neq N_{b}$;
1341: ({*}{*}) only zero modes were subtracted. For $x=0$ the \emph{RSS} result
1342: was obtained by a \emph{reciprocal space sum} using the analytical
1343: result.\cite{PAB04}}
1344: \end{figure}
1345:
1346: For $p=p_{\text{c}}$ we show in Fig.~\ref{cap:fssmz} the results obtained
1347: from three different approaches. The blue up-triangles are the results
1348: of our standard technique discussed in Sect.~\ref{sec:numeric}, \emph{i.e.},
1349: only lattices in which $N_{a}=N_{b}$ were considered and zero modes
1350: subtracted. The result labeled by violet down-triangles refers to
1351: a calculation in which the disordered realized lattices are not constrained
1352: to have $N_{a}=N_{b}$. The considerable difference between these two
1353: results is due to the presence of one {}``quasi-divergent'' low
1354: energy (nonzero) mode when $N_{a}\neq N_{b}$. That is, even though
1355: we subtract the zero energy Goldstone mode as discussed in
1356: Sect.~\ref{sec:numeric} for $N_{a}\neq N_{b}$, there is, in this situation,
1357: a low energy eigenstate that contributes in order $O(1)$ for $\delta m_{z}$,
1358: compared to the $O(1/N_{\text{c}})$ contributions of the others eigenstates.
1359: If the contribution of this mode is subtracted the result labeled
1360: by orange diamonds is obtained, which agrees well with the result
1361: of our standard technique (where the constrain $N_{a}=N_{b}$ is always
1362: used).
1363:
1364: To better understand the presence of this nonzero energy {}``quasi-divergent''
1365: mode when $N_{a}\neq N_{b}$, we have computed the contribution to
1366: $\delta m_{z}$ from the lower nonzero energy mode ($\delta m_{z}^{(1)}$),
1367: and the next one in energy ($\delta m_{z}^{(2)}$), constrained to
1368: lattices with $N_{a}-N_{b}=\pm1$. Figure \ref{cap:lm-honey} shows
1369: the behaviour of $\delta m_{z}^{(1)}$ (upper panel) and $\delta m_{z}^{(2)}$
1370: (lower panel) with the average cluster size $\bar{N}_{\text{c}}\propto L^{D}$.
1371: The $\delta m_{z}^{(2)}$ contribution decreases with $\bar{N}_{\text{c}}$,
1372: signaling the linear increase of the number of modes that contribute
1373: to $\delta m_{z}$. Instead, the contribution $\delta m_{z}^{(1)}$
1374: increases with $\bar{N}_{\text{c}}$, and will be of $O(1)$ in the thermodynamic
1375: limit. As already mentioned in Sect.~\ref{sec:numeric}, if $N_{a}$
1376: and $N_{b}$ are both of magnitude $10^{23}$, then, if $N_{a}-N_{b}=\pm1$,
1377: there will be, for any practical purpose, two Goldstone modes and
1378: not only one. This statement should always be true if
1379: $\left|N_{a}-N_{b}\right|\ll N_{a}\sim N_{b}$. The results presented in top
1380: panel of Fig.~\ref{cap:lm-honey} agree with this general picture.
1381: Furthermore, they imply that even for small sizes there is a mode, which will
1382: be identified with a Goldstone mode in the thermodynamic limit,
1383: that contributes {}``macroscopically'' to $\delta m_{z}$, though having a
1384: finite energy.%
1385: \begin{figure}
1386: \begin{center}\includegraphics[%
1387: clip,
1388: scale=0.75]{fig6-hsd.eps}\end{center}
1389:
1390: \caption{\label{cap:lm-honey}Contributions to $\delta m_{z}$ from: the lower
1391: nonzero energy mode (upper panel); the lower energy mode higher than
1392: the lower nonzero energy mode (lower panel). The average was taken
1393: over $10^{4}$ disordered honeycomb lattices, with $N_{a}-N_{b}=\pm1$,
1394: at $p=p_{\text{c}}$.}
1395: \end{figure}
1396:
1397: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1398: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1399: \subsection{Staggered magnetization}
1400:
1401: The results we found for the quantum mechanical factor $m_{\text{qm}}(x)$
1402: are summarized in Fig.~\ref{cap:mz} for the honeycomb lattice (panel
1403: (\textbf{a})) and for the square lattice (panel (\textbf{b})). Three
1404: different values of spin, $S=\frac{1}{2},1,\frac{3}{2}$, are shown.
1405:
1406: In the undiluted limit we obtain $\delta m_{z}(0)\approx0.258$
1407: for the honeycomb lattice, and $\delta m_{z}(0)\approx0.197$ for
1408: the square lattice. These results are in excellent agreement with quantum
1409: Monte Carlo results, namely, $\delta m_{z}(0)=0.2323(6)$ for the
1410: spin $1/2$ Heisenberg antiferromagnet in the honeycomb lattice (see
1411: Subsect.~\ref{sub:QMC}), and $\delta m_{z}(0)=0.1930(3)$ in the square
1412: lattice.\cite{San97}
1413:
1414: The effect of the classical factor $m_{\text{cl}}(x)$ (not shown) is only
1415: significant very close to $p_{\text{c}}$, where it vanishes with exponent
1416: $5/36$.\cite{SAperc} Thus, for $S>\frac{1}{2}$ there is a classically
1417: driven order disorder transition at $p_{\text{c}}$. For $S=\frac{1}{2}$
1418: linear spin wave theory predicts a \emph{quantum critical point} in
1419: both the honeycomb and square lattices to occur at $x^{*}=0.85(1)$
1420: and $x^{*}=0.98(1)$, respectively. Similar results for the square
1421: lattice were obtained in Ref.~\onlinecite{MNC04}, though the limited
1422: number of averages over disorder prevented the authors to distinguish
1423: $x^{*}$ from $x=1$.
1424:
1425: The predicted quantum critical point is absent in quantum Monte Carlo
1426: calculations, either in the honeycomb lattice or in the square
1427: lattice.\cite{San02} As already mentioned in Sect.~\ref{sec:hamilt},
1428: we should not expect the validity of spin wave approximation when
1429: $\delta m_{z}\sim S$, because inequalities \eqref{eq:ineqAb} and
1430: \eqref{eq:ineqBb} break down in this situation. This is precisely what
1431: happens when disorder increases for $S=\frac{1}{2}$.%
1432: \begin{figure}
1433: \begin{center}\includegraphics[%
1434: scale=0.75]{fig7-hsd.eps}\end{center}
1435:
1436: \caption{\label{cap:mz}(color online) Average quantum mechanical factor
1437: $m_{\text{qm}}(x)$ vs dilution $x=(1-p)/(1-p_{\text{c}})$ for different
1438: values of spin the $S$. Panel (\textbf{a}) shows the results for the
1439: honeycomb lattice and panel (\textbf{b}) for the square lattice.}
1440: \end{figure}
1441:
1442: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1443: \subsubsection*{Comparison with experimental results}
1444:
1445: Now we compare our results for the staggered magnetization in the
1446: spin wave approximation with available experimental measurements on
1447: Mn$_{p}$Zn$_{1-p}$PS$_{3}$ and Ba(Ni$_{p}$Mg$_{1-p}$)$_{2}$V$_{2}$O$_{8}$:
1448:
1449:
1450: \paragraph{Mn$_{p}$Zn$_{1-p}$PS$_{3}$}
1451:
1452: The layered compound MnPS$_{3}$ is a $S=5/2$ Heisenberg
1453: antiferromagnet.\cite{KSY83} This huge spin value suggests that the spin
1454: wave approximation should work well in this case. Indeed, the average
1455: magnetic moment on the Mn atoms was found to be $4.5(2)\,\mu_{B}$ at 3.5 K
1456: in the pure material,\cite{GSK+00} in excellent agreement with our spin wave
1457: result $m\approx4.48\,\mu_{B}$. The effect of dilution in the average
1458: magnetic moment of Mn$^{2+}$ ions is presented in
1459: Fig.~\ref{cap:m-p.sw+exp.2S-5}. Neutron diffraction results on
1460: Mn$_{p}$Zn$_{1-p}$PS$_{3}$ are shown as grey circles,\cite{GSK+00}
1461: and the red squares are the theoretical results within the linear
1462: spin wave approximation. To go beyond $p_{\text{c}}$ (the first-nearest
1463: neighbor percolation threshold) we would have to take into account
1464: second- and third-nearest neighbor couplings in Hamiltonian \eqref{eq:hamilt1}.
1465: Nevertheless, the effect of dilution for $p\leq p_{\text{c}}$ is already
1466: well described by the first-nearest neighbor model. Furthermore,
1467: the agreement between experimental and theoretical results even at
1468: $p=p_{\text{c}}$, indicates that the primarily effect of second- and
1469: third-nearest neighbor interactions is classical. That is, the existence of one
1470: largest connected cluster with a finite fraction of spins in the thermodynamic
1471: limit is guaranteed by this couplings for $p>p_{\text{c}}$, but the quantum
1472: correction to the staggered magnetization density is determined by
1473: the smaller first-nearest neighbors clusters belonging to this larger
1474: one, at least for $p\gtrsim p_{\text{c}}$. Further investigations are needed
1475: to clarify whether this is the correct picture.\cite{CP05}%
1476: \begin{figure}
1477: \begin{center}\includegraphics[%
1478: scale=0.75]{fig8-hsd.eps}\end{center}
1479:
1480: \caption{\label{cap:m-p.sw+exp.2S-5}(color online) Average magnetic moment
1481: per magnetic site as function of dilution $p$. The linear spin wave
1482: result for the $S=5/2$ Heisenberg antiferromagnet in the honeycomb
1483: lattice (red squares) is compared with neutron scattering data on
1484: Mn$_{p}$Zn$_{1-p}$PS$_{3}$ from Ref.~\onlinecite{GSK+00} (grey
1485: circles).}
1486: \end{figure}
1487:
1488:
1489: \paragraph{Ba(Ni$_{p}$Mg$_{1-p}$)$_{2}$V$_{2}$O$_{8}$}
1490:
1491: The layered compound BaNi$_{2}$V$_{2}$O$_{8}$ is a spin $S=1$
1492: antiferromagnet in a honeycomb lattice. Neutron diffraction experiments
1493: have found, in the pure case, an average magnetic moment of $1.55(4)\,\mu_{B}$
1494: for Ni at 8 K,\cite{RHL+02} which is in good agreement with the spin
1495: wave result $m\approx1.48\,\mu_{B}$. To our knowledge, the magnetic
1496: moment has not yet been measured for the diluted compound. Nevertheless,
1497: the available magnetic susceptibility measurements on
1498: Ba(Ni$_{p}$Mg$_{1-p}$)$_{2}$V$_{2}$O$_{8}$ for dilutions in the range
1499: $0.84\leq p\leq1$, show that the N\'{e}el temperature is strongly dependent
1500: on the amount of dilution.\cite{RHL+02} For the highest diluted sample
1501: ($p=0.84$) a reduction of almost 70\% relative to the undiluted N\'{e}el
1502: temperature was found. It would be interesting to know whether the suppression
1503: of antiferromagnetic LRO by nonmagnetic impurities will occur at the classical
1504: percolation transition $p_{\text{c}}\simeq0.7$, as predicted in our
1505: calculations.
1506:
1507: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1508: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1509: \subsection{N\'{e}el Temperature}
1510:
1511: The N\'{e}el temperature of both Mn$_{p}$Zn$_{1-p}$PS$_{3}$ and
1512: Ba(Ni$_{p}$Mg$_{1-p}$)$_{2}$V$_{2}$O$_{8}$ shows a linear suppression
1513: with increasing dilution $1-p$,\cite{GSK+00,RHL+02}
1514: a feature that is also seen in (quasi-2D) diluted Heisenberg antiferromagnets
1515: with square lattice.\cite{CRS97,HKP+99,TFH+00}
1516:
1517: Within the linear spin wave theory developed in Secs. \ref{sec:hamilt}
1518: and \ref{sec:numeric} for diluted antiferromagnetic systems the finite
1519: temperature staggered magnetization is given by
1520: \begin{equation} \label{eq:magzT}
1521: \begin{split}
1522: M_{z}^{\text{stagg}}(T) & =
1523: \left\langle \sum_{i\in A}S_{i}^{a,z}-\sum_{i\in B}S_{i}^{b,z}\right\rangle \\
1524: = & N_{\text{c}}\bigl(S-\delta m_{z}-\delta m_{z}^{T}(T)\bigr),
1525: \end{split}
1526: \end{equation}
1527: where $\delta m_{z}$ is the zero-temperature correction to the staggered
1528: magnetization defined in Eq.~\eqref{eq:deltamz}, and $\delta m_{z}^{T}(T)$
1529: is the thermal correction
1530: \begin{equation}
1531: \delta m_{z}^{T}(T)=\sum_{n=1}^{N_{a}}\delta m_{z}^{(n,\alpha)}
1532: n_{B}(\omega_{n}^{(\alpha)})+\sum_{n=1}^{N_{b}}\delta m_{z}^{(n,\beta)}
1533: n_{B}(\omega_{n}^{(\beta)}),
1534: \label{eq:deltamzT}
1535: \end{equation}
1536: with generalized $\delta m_{z}^{(n,\alpha)}$ and $\delta m_{z}^{(n,\beta)}$,
1537: \begin{align}
1538: \delta m_{z}^{(n,\alpha)} & = \frac{1}{N_{\text{c}}}
1539: \Biggl(\sum_{i\in A}|u_{ni}|^{2}+\sum_{i\in B}|v_{ni}|^{2}\Biggr),
1540: \label{eq:deltamzalphT}\\
1541: \delta m_{z}^{(n,\beta)} & = \frac{1}{N_{\text{c}}}
1542: \Biggl(\sum_{i\in A}|w_{ni}|^{2}+\sum_{i\in B}|x_{ni}|^{2}\Biggr),
1543: \label{eq:deltamzbetT}
1544: \end{align}
1545: and $n_{B}(\omega)=(e^{\omega/k_{B}T}-1)^{-1}$ is the Bose distribution
1546: function. In the thermodynamic limit the averaged over disorder staggered
1547: magnetization density can be expressed as
1548: \begin{equation}
1549: m_{\text{av}}(p,T,L\rightarrow\infty) = m_{\text{cl}}m_{\text{qm}}(T),
1550: \label{eq:mavLinftyT}
1551: \end{equation}
1552: where $m_{\text{cl}}$ is the classical factor defined in Eq.~\eqref{eq:mcl},
1553: and $m_{\text{qm}}(T)$ is the temperature dependent quantum mechanical factor,
1554: \begin{equation}
1555: m_{\text{qm}}(T) =
1556: S-\delta m_{z}^{\infty}-\delta m_{z}^{T,L\rightarrow\infty}(T).
1557: \label{eq:mqmT}
1558: \end{equation}
1559:
1560: In the undiluted case the thermal correction $\delta m_{z}^{T}(p=1,T)$
1561: can be expressed as
1562: \begin{equation}
1563: \delta m_{z}^{T,L}(p=1,T)=\frac{1}{N_{a}+N_{b}}\sum_{\mathbf{k}}
1564: \frac{h_{a}}{\sqrt{h_{a}^{2}-\left|\phi_{\mathbf{k}}\right|^{2}}}
1565: n_{B}(\omega_{\mathbf{k}}),
1566: \label{eq:deltamzTnd}
1567: \end{equation}
1568: with $\omega_{\mathbf{k}}$ as in Eq.~\eqref{eq:Ek}, and $\phi_{\mathbf{k}}$
1569: given by Eq.~\eqref{eq:phik}. The summation in $\mathbf{k}$ is done
1570: in the first Brillouin zone of sublattice $A$ or $B$, and can be
1571: replaced by an integration when $L\rightarrow\infty$. When $h_{a}=1$
1572: the spin wave dispersion behaves as $\omega_{\mathbf{k}}\propto k$
1573: in the long wave length limit, similarly to the square lattice case.
1574: As a consequence the thermal correction to the staggered magnetization
1575: develops a logarithmic divergence, which signals the well known suppression
1576: of LRO at $T>0$ in the 2D isotropic Heisenberg model.
1577:
1578: Therefore, if LRO is present up to $T_{N}\neq0$, either a magnetic
1579: anisotropy $h_{a}$ or a finite interplanar exchange $J_{\perp}$
1580: (or both) must be present. If the former is the dominant effect $T_{N}$
1581: can be calculated using the mean-field like equation
1582: \footnote{The N\'eel temperature determined from Eq.~\eqref{eq:Tmfgap}
1583: for isotropic Heisenberg systems is known to
1584: be overestimated. In order to correct for it a self consistent solution
1585: of this equation is used, where $S$ is replaced by $m_{\rm av}$. This
1586: procedure, very simple to implement in the undiluted case, since there
1587: is an analytical expression available for the magnon spectrum, is much
1588: more difficult in our case. We postpone the discussion of this aspect to
1589: a latter publication.}
1590: \begin{equation}
1591: m_{\text{qm}}(T_{N})=0.
1592: \label{eq:Tmfgap}
1593: \end{equation}
1594: In the latter the transition should occur when the interplanar coupling
1595: is strong enough to stabilize the LRO in comparison with thermal fluctuations:
1596: \begin{equation}
1597: J_{\perp}m_{\text{qm}}^{2}(p,T=0)\frac{\xi^{2}(p,T_{N})}{A/2}
1598: \approx k_{B}T_{N},
1599: \label{eq:TmfJperp}
1600: \end{equation}
1601: The parameter $\xi(p,T)$ is the inplane correlation length, which
1602: characterizes the spin fluctuations of a layered system in a paramagnetic
1603: phase. The area of a hexagon of side $c$ is given by $A=c^{2}3\sqrt{3}/2$.
1604: The correlation length can be calculated in the context of the modified
1605: spin wave theory,\cite{TAK89} and in the non-diluted ($p=1$) case
1606: it is exponentially divergent with $1/T$ as $T\rightarrow0$. The
1607: mean field picture which leads to Eq.~\eqref{eq:TmfJperp} was proposed
1608: in Ref.~\onlinecite{CCN02}, and gives a good description of the variation
1609: of $T_{N}(p)/T_{N}(0)$ with dilution $1-p$ in a variety of layered
1610: compounds with square lattice.
1611:
1612: In the case of MnPS$_{3}$ a small gap of magnitude $\Delta E=0.5$~meV
1613: was found in the spin wave energy at the the Brillouin zone
1614: center.\cite{WRL+98} This energy gap can be explained by either a
1615: single-ion anisotropy or a dipole coupling, being modeled here by a small
1616: magnetic anisotropy $h_{a}>1$. From the spin wave dispersion \eqref{eq:Ek}
1617: it is found that $h_{a}\approx1.004$ is needed to obtain $\Delta E=0.5$ meV
1618: (a nearest-neighbor exchange of magnitude $J=0.8$ meV was used\cite{WRL+98}).
1619: We remark that such a small magnetic anisotropy has no effect in the
1620: conclusions we have made so far based in the isotropic Heisenberg
1621: model ($h_{a}=1$). As an example, the the average magnetic moment
1622: on the Mn atoms given by spin-wave theory is $m\approx4.48\,\mu_{B}$
1623: for $h_{a}=1$ and $m\approx4.55\,\mu_{B}$ for $h_{a}=1.004$, both
1624: in excellent agreement with the experimental value $4.5(2)\,\mu_{B}$
1625: at 3.5~K.\cite{GSK+00} Inserting this value of $h_{a}$ into Eq.
1626: (\ref{eq:deltamzTnd}) we obtain $T_{N}\approx70$~K as a solution
1627: of Eq.~(\ref{eq:Tmfgap}), in agreement with the measured value
1628: $T_{N}=78$~K.\cite{KSY83}
1629:
1630: Nevertheless a finite interplanar exchange of magnitude
1631: $J_{\perp}=0.0019(2)$~meV is also present in the MnPS$_{3}$
1632: compound.\cite{WRL+98}
1633: With $\xi(p\!=\!1,T_{N}\!=\!78~\text{K})=27.5~\text{\AA}$
1634: measured by neutron scattering,\cite{RWB00} and $c=3.5~\text{\AA}$,
1635: \cite{okuda-mpolcm1986} we obtain from the mean field equation
1636: \eqref{eq:TmfJperp} $T_{N}\approx6$~K.
1637: This small value of $T_{N}$ is an indication that the effect of the
1638: interplanar coupling is not as important as the magnetic anisotropy
1639: in stabilizing the LRO. Therefore we use Eq.~\eqref{eq:Tmfgap} to
1640: study the effect of dilution on $T_{N}(p)$. The thermal correction
1641: $\delta m_{z}^{T}$ defined by Eq.~\eqref{eq:deltamzT} is computed via
1642: recursion method (see Subsect.~\ref{sub:DOSrm}), noting that it can be
1643: expressed as
1644: \begin{equation}
1645: \delta m_{z}^{T}(T)=\int_{0}^{\infty}dEn_{B}(E)K(E),
1646: \label{eq:deltamzTrm}
1647: \end{equation}
1648: where the kernel $K(E)$ is given by
1649: \begin{multline} \label{eq:kernel}
1650: K(E) = -\frac{1}{N_{\text{c}}}\frac{1}{\pi}\text{Im}
1651: \Biggl[\sum_{i\in A}G_{ii}^{aa}(E+i0^{+})\\
1652: +\sum_{i\in B}G_{ii}^{bb}(E+i0^{+}) + \sum_{i\in A}G_{ii}^{aa}(-E+i0^{+})\\
1653: +\sum_{i\in B}G_{ii}^{bb}(-E+i0^{+})\Bigr].
1654: \end{multline}
1655: It is worth mentioning that with the recursion method $\delta m_{z}^{T}$
1656: can be computed with the same precision (limited by the linear size
1657: $L=128$ of the sample) from the undiluted limit $p=1$ to the percolation
1658: threshold $p=p_{\text{c}}$.
1659:
1660: The result of numerically solving Eq.~\eqref{eq:Tmfgap}---with
1661: $\delta m_{z}^{T}$ computed by applying the recursion method to systems
1662: with $L=128$ and averaging over 200 to 400 disorder realizations---is shown in
1663: Fig.~\ref{cap:TN2S-5}.%
1664: \begin{figure}
1665: \begin{center}\includegraphics[%
1666: clip,
1667: scale=0.75]{fig9-hsd.eps}\end{center}
1668:
1669: \caption{\label{cap:TN2S-5}(color online) $T_{N}(p)/T_{N}(0)$ vs $p$ for
1670: $S=5/2$. Shown are the results obtained by numerically solving Eq.
1671: \eqref{eq:Tmfgap} with $\delta m_{z}^{T}$ computed applying the
1672: recursion method to systems with $L=128$ and averaging over 200 to
1673: 400 disorder realizations (squares), the mean-field result of Eq.
1674: (\ref{eq:TMFp}) (diamonds), and experimental results on
1675: Mn$_{p}$Zn$_{1-p}$PS$_{3}$ from Ref.~\onlinecite{GSK+00} (circles).}
1676: \end{figure}
1677: Also shown are the results of magnetometry measurements on
1678: Mn$_{p}$Zn$_{1-p}$PS$_{3}$ from Ref. \onlinecite{GSK+00}. The difference
1679: between the theoretical results and experimental values suggests that in
1680: opposition to the magnetic moment at zero temperature
1681: (see Fig. \ref{cap:m-p.sw+exp.2S-5}) the effect of second- and
1682: third-nearest-neighbor couplings should be included to obtain a quantitatively
1683: correct N\'{e}el temperature as dilution is increased. An estimation of
1684: $T_{N}(p)/T_{N}(1)$ can as well be obtained by standard mean-field theory,
1685: $T_{N}^{MF}=\frac{2}{3}JzS(S+1)$.\cite{MJintheorPhys}
1686: Replacing $S$ by the zero temperature staggered magnetization density
1687: $m_{av}(p)$ defined in Eq.~(\ref{eq:mavLinfty}), and assuming that
1688: the coordination number decreases linearly with dilution, $z\propto p$,
1689: the ratio $T_{N}(p)/T_{N}(1)$ is given by
1690: \begin{equation}
1691: \frac{T_{N}(p)}{T_{N}(1)} = p\, m_{av}(p)[m_{av}(p)+1].
1692: \label{eq:TMFp}
1693: \end{equation}
1694: In Fig.~\ref{cap:TN2S-5} we show as diamonds the results of Eq.
1695: (\ref{eq:TMFp}). Although this result reproduces the correct dependence
1696: on $p$, it should be stressed that as a mean-field approximation
1697: the absolute value of $T_{N}(p)$ is overestimated.
1698:
1699: The effect of dilution on the N\'{e}el temperature of
1700: Ba(Ni$_{p}$Mg$_{1-p}$)$_{2}$V$_{2}$O$_{8}$ was studied by Rogado
1701: \emph{et al.} for dilutions in the range $0.84\leq p\leq1$.\cite{RHL+02}
1702: The few experimental results concerning the magnetic properties of
1703: BaNi$_{2}$V$_{2}$O$_{8}$ are insufficient to undoubtedly determine
1704: the model which better describes the magnetic behaviour of this compound.
1705: Although electron-spin resonance measurements seem to be well fitted by a
1706: weakly anisotropic Heisenberg model with easy-plane symmetry (XY), \emph{i.e.},
1707: $h_{a}\lesssim1$ in Hamiltonian \eqref{eq:haf},
1708: the same results can as well be explained with the isotropic limit
1709: of this model.\cite{HvN+03} Further experiments would be valuable
1710: in determining the nature of the LRO observed in this compound, in
1711: particular inelastic neutron scattering from which the spin wave dispersion
1712: can be measured. Here we assume that a small gap is present at the
1713: Brillouin zone center, and that it can be modeled by a small uniaxial
1714: interaction anisotropy, \emph{i.e.}, $h_{a}\gtrsim1$ in Hamiltonian
1715: \eqref{eq:haf}. In particular $h_{a}-1\approx10^{-4}$ is needed to get
1716: $T_{N}\approx50$~K in the undiluted case (a nearest-neighbor exchange of
1717: magnitude $J\approx4$~meV was used).\cite{RHL+02}
1718:
1719: The $T_{N}(p)/T_{N}(1)$ vs $p$ result obtained by numerically solving
1720: Eq.~(\ref{eq:Tmfgap}) for $S=1$, with $\delta m_{z}^{T}$ computed
1721: applying the recursion method to systems with $L=128$ and averaging
1722: over 200 to 400 disorder realizations is shown in Fig.~\ref{cap:TNS-1}
1723: (squares).%
1724: \begin{figure}
1725: \begin{center}\includegraphics[%
1726: scale=0.75]{fig10-hsd.eps}\end{center}
1727:
1728: \caption{\label{cap:TNS-1}(color online) $T_{N}(p)/T_{N}(0)$ vs $p$ for
1729: $S=1$. Shown are the results obtained by numerically solving
1730: Eq.~\eqref{eq:Tmfgap} with $\delta m_{z}^{T}$ computed applying the
1731: recursion method to systems with $L=128$ and averaging over 200 to 400
1732: disorder realizations (squares), the mean-field result of Eq.~(\ref{eq:TMFp})
1733: (diamonds), and experimental results on
1734: Ba(Ni$_{p}$Mg$_{1-p}$)$_{2}$V$_{2}$O$_{8}$ from Ref.~\onlinecite{RHL+02}
1735: (circles).}
1736: \end{figure}
1737: Also shown are the mean-field result of Eq.~(\ref{eq:TMFp}) (diamonds)
1738: and results of magnetic susceptibility data for
1739: Ba(Ni$_{p}$Mg$_{1-p}$)$_{2}$V$_{2}$O$_{8}$ (circles) (Ref.
1740: \onlinecite{RHL+02}). The disagreement between the
1741: mean-field result (Eq.~(\ref{eq:TMFp})) and experimental values can
1742: be attributed to the small spin $S=1$ value, which means higher quantum
1743: fluctuations and less mean-field like behaviour. The theoretical result
1744: (squares) and the experimental values are in reasonable agreement,
1745: though it seems to worsen as dilution increases. It should be noted
1746: that the spin-wave theory for layered materials is not really adequate
1747: at $T\sim T_{N}$, and when it is applied to the mean-field like
1748: Eq.~(\ref{eq:Tmfgap}) it tends to overestimate the absolute value of
1749: the N\'{e}el temperature.\cite{IKK99}
1750:
1751: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1752: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1753: \subsection{Density of states}
1754:
1755: The effect of dilution has a strong impact on the DOS of the system.
1756: Since the momentum is no-longer a well defined quantum number the
1757: spin waves acquire a finite lifetime.\cite{CCN02} As a consequence,
1758: the basis that diagonalizes the problem has a very different energy
1759: spectrum, which implies a different DOS.
1760:
1761: We have calculated the DOS of the antiferromagnetic Heisenberg model
1762: in the linear spin wave approximation for the honeycomb and square
1763: lattices in the presence of dilution. The recursion method briefly
1764: discussed in Subsect. \ref{sub:DOSrm} was used to study the variation
1765: of the DOS with dilution. The method is valid from the undiluted
1766: $p=1$ limit to the percolation threshold $p_{c}$, and enables the
1767: access to the whole energy spectrum. The precision limit is set by
1768: the linear size $L$ of the system, which we fix here to $L=128$
1769: both in the honeycomb and square lattices.
1770:
1771: In Fig.~\ref{cap:DOSsquare} we show the square lattice DOS at four
1772: different values of dilution $x$.%
1773: \begin{figure}
1774: \begin{center}\includegraphics[%
1775: clip,
1776: scale=0.75]{fig11-hsd.eps}\end{center}
1777:
1778: \caption{\label{cap:DOSsquare}(color online) DOS of the Heisenberg
1779: antiferromagnetic model in the linear spin wave approximation for the
1780: square lattice. An energy mesh with spacing $0.01$ in units of $JS$
1781: was used. This results were obtained applying the recursion method to
1782: systems with $L=128$, and averaging over 200 to 400 disorder realizations. The
1783: dotted line is the clean limit DOS.}
1784: \end{figure}
1785: The depletion of the high energy part of the DOS in favor of low
1786: energy modes is clearly seen as dilution is increased, in agreement
1787: with the results obtained by exact diagonalize smaller systems.\cite{MNC04}
1788: The two structures visible at around $E/JS=2$ and $3$, which Mucciolo
1789: \emph{et al.}\cite{MNC04} associated with the breaking of the clean-limit
1790: magnon branch into three distinct but broad branches, are also evident.
1791:
1792: The DOS for the honeycomb lattice is shown in Fig. \ref{cap:DOShoney}.
1793: A decrease in the density of high-frequency states and the proportional
1794: increase in the density of low-frequency ones is also clear as dilution
1795: increases.%
1796: \begin{figure}
1797: \begin{center}\includegraphics[%
1798: clip,
1799: scale=0.75]{fig12-hsd.eps}\end{center}
1800:
1801: \caption{\label{cap:DOShoney}(color online) DOS of the Heisenberg
1802: antiferromagnetic model in the linear spin wave approximation for
1803: the honeycomb lattice. An energy mesh with spacing $0.01$ in units of
1804: $JS$ was used. This results were obtained applying the recursion method
1805: to systems with $L=128$, and averaging over 200 to 400 disorder realizations.
1806: The dotted line is the clean limit DOS.}
1807: \end{figure}
1808: This feature can then be viewed as a general effect of the presence
1809: of dilution. Structures as those observed in the square lattice case,
1810: just below $E/JS=2$ and 3, are not so easily identified. Nevertheless,
1811: a feature of this kind seems to be present just below $E/JS=2$. To
1812: determine whether or not it can be associated to the presence of fractons,
1813: as in the square lattice case,\cite{MNC04} a more detailed study
1814: is needed, such as the calculation of the dynamical structure factor
1815: in the diluted honeycomb lattice.
1816:
1817: The effect of moving spectral weight from the top of the band to lower
1818: energies due to dilution is accompanied by the appearance of a set
1819: of peaks, starting to develop in the high-frequency part of the spectrum
1820: for small dilution and extending to the entire band as dilution increases.
1821: There is, however, a particular peak that deserves special attention.
1822: This peak can be seen very close to the bottom of the band ($E=0$)
1823: for $x\geq0.8$ both in the honeycomb and square lattice DOS. Figure
1824: \ref{cap:DOSlowener} is a zoom of the DOS close to $E=0$ at $x=x_{c}$.%
1825: \begin{figure}
1826: \begin{center}\includegraphics[%
1827: clip,
1828: scale=0.75]{fig13-hsd.eps}\end{center}
1829:
1830: \caption{\label{cap:DOSlowener}(color online) Low energy behaviour of the
1831: DOS of the Heisenberg antiferromagnetic model in the linear spin wave
1832: approximation for the honeycomb (left) and square (right) lattices
1833: at $x=x_{c}$. An energy mesh with spacing $5\times10^{-4}$ in units
1834: of $JS$ was used. This results were obtained applying the recursion
1835: method to systems with $L=128$, and averaging over 800 disorder realizations.}
1836: \end{figure}
1837: Being present both in the honeycomb and square lattices, though a
1838: bit stronger in the former, this peak seems to be a general feature
1839: associated with dilution. In fact, it is closely related to the finiteness
1840: of the quantum corrections to the staggered magnetization at zero
1841: temperature.
1842:
1843: As shown by Mucciolo \emph{et al.}\cite{MNC04}, the finiteness of
1844: the quantum fluctuations reduces to the problem of the convergence
1845: of the integral $\int_{0}^{E_{max}}dE\rho(E)E^{-1}$. In
1846: Fig.~\ref{cap:DOSlowener} we show a polynomial fit to the low-energy behaviour
1847: of the DOS (red line in the left side of the peak). Although it should be
1848: seen as guide to the eyes, we can undoubtedly say that in the low-energy limit
1849: the DOS behaves as $\rho(E)\propto E^{\alpha}$ with $\alpha>1$,
1850: and thus the above mentioned integral is convergent. This result is
1851: consistent with the existence of an upper bound for the quantum fluctuations
1852: in any model with a classically ordered ground state whose Hamiltonian
1853: can be mapped onto that of a system of coupled harmonic oscillators,
1854: argued by Mucciolo \emph{et al.}.\cite{MNC04} This result also agrees
1855: with the FSS results presented in Subsect. \ref{sub:Finite-size-scaling},
1856: where we found finite values for $\delta m_{z}^{\infty}(x)$. And
1857: the fact that $\delta m_{z}^{\infty}(x_{c})>1/2$ can be attributed
1858: to the bad-behaviour of the spin-wave approximation when $\delta m_{z}\sim S$,
1859: as will be shown in the next section.
1860:
1861: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1862: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1863: \subsection{Quantum Monte Carlo results for $S=1/2$}
1864: \label{sub:QMC}
1865:
1866: We have performed a Monte Carlo study of the $S=1/2$ quantum Heisenberg
1867: antiferromagnet on the site-diluted honeycomb lattice using Stochastic
1868: Series Expansion (SSE).~\cite{San97,San99} Unlike the spin wave
1869: approach described in Sects.~\ref{sec:hamilt} and \ref{sec:numeric}---which
1870: should be understood as an expansion in the relative reduction of
1871: the staggered moment $\delta m_{z}/S$---this technique is exact (up
1872: to statistical uncertainties) and well-behaved even when $\delta m_{z}\sim S$.
1873: In particular, the SSE Monte Carlo can access the the small-$S$,
1874: near-percolation regime where the spin wave calculation becomes unreliable.
1875:
1876: We have closely followed the procedure outlined in Ref.~\onlinecite{San02},
1877: which treats the site dilution problem on the square lattice. To accelerate
1878: convergence, we have taken advantage of the $\beta$-doubling scheme
1879: described therein: 100 equlibration and 200 sampling sweeps are performed
1880: at each temperature with the resulting configuration (an $M$-element
1881: operator list $S_{M}=[a_{1},b_{1}],\ldots,[a_{M},b_{M}]$) used to
1882: generate a high-probability initial configuration at the next lowest
1883: temperature ($S_{2M}=[a_{1},b_{1}],\ldots,[a_{M},b_{M}],[a_{M},b_{M}],\ldots,[a_{1},b_{1}]$)
1884: according to the cooling schedule $\beta=2,4,8,\ldots,2048,4096$.
1885:
1886: A refinement to previous work is that we extrapolate the staggered
1887: magnetization to the thermodynamic limit
1888: using \emph{two} different quantities:
1889: \begin{subequations} \label{EQ:minfty}
1890: \begin{align} \label{EQ:minftyA}
1891: m_{\text{qm}} &= \lim_{L\rightarrow\infty} \biggl\langle \frac{2}{N_{\text{c}}}\Bigl\lvert \,
1892: \hat{M}^{\text{stagg}}_z \,\Bigr\rvert \biggr\rangle_{\!L,x},\\ \label{EQ:minftyB}
1893: m_{\text{qm}}^2 &= \lim_{L\rightarrow\infty} \biggl\langle \frac{3}{N_{\text{c}}^2}\Bigl(
1894: \hat{M}^{\text{stagg}}_z \Bigr)^2 \biggr\rangle_{\!L,x}.
1895: \end{align}
1896: \end{subequations}
1897: Here, $\hat{M}^{\text{stagg}}_z = \sum_{i\in A} \hat{S}_i^z - \sum_{i\in B} \hat{S}_i^z$
1898: is the $z$-projected staggered magetization and $m_{\text{qm}}$ is the quantum mechanical factor introduced
1899: in Subsect.~\ref{sub:Finite-size-scaling}.
1900: The notation $\langle\,\cdot\,\rangle_{L,x}$ represents an ensemble average
1901: over the quantum states of the system and over all configurations
1902: of the size-$L$ lattice with dilution $x$. The site indices in $\hat{M}^{\text{stagg}}_z$ are
1903: understood to range over only the largest connected cluster.
1904:
1905: Equation~\eqref{EQ:minftyA}, being linear, is analogous to the quantity
1906: $S-\delta m_{z}$ computed via spin wave theory. Equation~\eqref{EQ:minftyB}
1907: is essentially a structure factor and equivalent to Eq.~(10) of Ref.~\onlinecite{San02}.
1908: The factors 2 and 3 in Eqs.~\eqref{EQ:minfty} are a consequence
1909: of the rotational invariance of the ground state. Their particular
1910: values follow from the averages $\int\! d\hat{\Omega}\,\lvert\hat{\Omega}\cdot\hat{z}\rvert=4\pi/2$
1911: and $\int\! d\hat{\Omega}\,\bigl(\hat{\Omega}\cdot\hat{z}\bigr)^{2}=4\pi/3$
1912: where $\hat{\Omega}$ is a vector ranging over the unit sphere. (Such
1913: geometric factors are irrelevant to the spin wave case; there the
1914: ground state is symmetry-broken by explicit construction.)
1915:
1916: As in Ref.~\onlinecite{San02}, we use the straight-forward generalization
1917: of the finite-size scaling form for the clean system,\cite{HUS88}
1918: \begin{subequations} \label{EQ:fssansatz}
1919: \begin{align}
1920: \label{EQ:fssansatzA}
1921: \biggl\langle \frac{2}{N_{\text{c}}}\Bigl\lvert \,
1922: \hat{M}^{\text{stagg}}_z \,\Bigr\rvert \biggr\rangle_{\!\!L,x}^{2} &
1923: \!= m_{\text{qm}}^2 + \frac{a_1}{\sqrt{\bar{N}_{\text{c}}}} + \frac{a_2}{\bar{N}_{\text{c}}} + \cdots, \\
1924: \label{EQ:fssansatzB}
1925: \biggl\langle \frac{3}{N_{\text{c}}^2}\Bigl(
1926: \hat{M}^{\text{stagg}}_z \Bigr)^2 \biggr\rangle_{\!\!L,x} &
1927: \!= m_{\text{qm}}^2 + \frac{b_1}{\sqrt{\bar{N}_{\text{c}}}} + \frac{b_2}{\bar{N}_{\text{c}}} + \cdots.
1928: \end{align}
1929: \end{subequations}
1930: [As the discussion in Subsect.~\ref{subsec:clusterstat} makes clear, this
1931: converges to $L^{-D/2}$ powerlaw behavior at large $L$, as in Eq.~\eqref{eq:scaling}.]
1932: Numerical measurements of the two quantities on the left-hand side
1933: of Eqs.~\eqref{EQ:fssansatzA} and \eqref{EQ:fssansatzB} may be
1934: fit to the corresponding functions on the right-hand side either simulaneously---with
1935: parameters $m_{\text{qm}}$, $\{a_{i}\}$, $\{b_{i}\}$---or separately---with parameters $m_{\text{qm}}$,
1936: $\{a_{i}\}$ and $m_{\text{qm}}'$, $\{b_{i}\}$. Verifying that $m_{\text{qm}}\approx m_{\text{qm}}'$
1937: serves as a consistency check.
1938: \begin{figure}
1939: \begin{center}\includegraphics[%
1940: scale=0.85]{fig14-hsd.eps}\end{center}
1941: \caption{\label{FIG:magundiluted} The staggered magnetization of the undiluted
1942: honeycomb lattice ($x=0$, $N_{\text{c}}=N=2L^{2}$) is extrapolated
1943: to the thermodynamic limit following Eqs.~\eqref{EQ:fssansatzA}
1944: and \eqref{EQ:fssansatzB}. A simultaneous fit of the two data sets
1945: yields the value $m_{\text{av}}(L\to\infty)=0.2677(6)$.}
1946: \end{figure}
1947:
1948:
1949: In the case of the \emph{undiluted} honeycomb lattice
1950: (for which $m_{\text{av}}(L\to\infty)\equiv m_{\text{qm}}$),
1951: we have simulated lattices up to linear size $L=32$
1952: (\emph{i.e.}, up to $2\times32^{2}=2048$ sites). Observables were
1953: computed using a bootstrap analysis~\cite{ET93} of 150 bins of $10^{5}$
1954: samples each ($1.5\times10^{6}$ total Monte Carlo sweeps). Best fits
1955: to the data, shown in Fig.~\ref{FIG:magundiluted}, give the thermodynamic
1956: limit $m_{\text{av}}(L\rightarrow\infty)=0.2677(6)$.
1957: This is somewhat smaller than the square lattice value $m_{\text{av}}(L\rightarrow\infty)=0.3070(3)$,\cite{San02}
1958: a reduction that reflects the larger quantum fluctuations on the less
1959: meanfield-like honeycomb lattice.
1960:
1961: Note that our value of the staggered magnetization is larger than
1962: (but consistent with) an earlier Monte Carlo measurement due to Reger
1963: \emph{et al}.~\cite{RRY89} (within 1.6 standard deviations). It
1964: is also, we believe, considerably more accurate. The Reger group's
1965: value of $m_{\text{av}}(L\rightarrow\infty)=0.22(3)$
1966: was computed by extrapolating relatively large Trotter errors ($0.1<\Delta\tau<0.2$)
1967: to $\Delta\tau\rightarrow0$ and small systems sizes ($4<L<8$) to
1968: $L\rightarrow\infty$. Moveover, their analysis supposes that the
1969: inverse temperature $\beta=10$ is sufficiently cold to extract the
1970: ground state properties of the system, which is very likely incorrect.~\cite{San02}%
1971: \begin{figure}
1972: \begin{center}\includegraphics[%
1973: scale=0.85]{fig15-hsd.eps}\end{center}
1974:
1975:
1976: \caption{\label{FIG:magdiluted} (color online) The main plot shows an extrapolation
1977: to the thermodynamic limit of twice the $z$-projected staggered magnetization
1978: for various dilution levels $x$ (as indicated by the symbols in the
1979: upper-left legend). The lines drawn through the data points represent
1980: a global fit to Eqs.~\eqref{EQ:fssansatz} in which $m_{\text{qm}}(x),a_{1}(x),b_{1}(x),\ldots$
1981: are treated as powerseries in $x$ and varied. The resulting function
1982: $m_{\text{qm}}(x)$ appears as the solid
1983: (pink) line in the figure inset alongside Monte Carlo results for
1984: the square lattice (from Ref.~\onlinecite{San02}) and spinwave results
1985: for the honeycomb lattice. The (red) errorbars indicate the values
1986: of $m_{\text{qm}}$ extrapolated from each
1987: fixed-$x$ dataset taken individually. }
1988: \end{figure}
1989:
1990:
1991: For the \emph{diluted} honeycomb lattice, we computed the staggered
1992: magnetization as an average over $10^{5}$ randomly-generated disorder
1993: realizations. Simulations of system sizes up to $\bar{N}_{\text{c}} \approx 2000$ were extrapolated
1994: to the thermodynamic limit, as shown in Fig.~\ref{FIG:magdiluted}.
1995: The figure inset illustrates the dependence of $m_{\text{qm}}$ on dilution.
1996:
1997: In contrast to the spinwave prediction, we find that LRO
1998: persists right up to the classical percolation threshold. The magnitude
1999: of the staggered magnetization decreases with dilution but does not
2000: vanish: $m_{\text{qm}}=0.139(6)$ at $x=1$, which represents a roughly $50\%$ reduction in magnetic
2001: moment over the undiluted ($x=0$) lattice. This is comparable to
2002: the effect seen in the square lattice where $m_{\text{qm}}(0)=0.3070(3)$
2003: falls to $m_{\text{qm}}(1)=0.150(2)$.
2004:
2005: We observe that the square- and honeycomb-lattice values of $m_{\text{qm}}$
2006: are remarkably close in the vicinity of $x=1$. The likely
2007: explanation is that the percolating clusters---retaining little of
2008: the structure of their undiluted parent lattice---are themselves quite
2009: similar. Both have fractal dimension $D=91/48$ and a similar nearest
2010: neighbour count: with increasing site dilution, the average coordination
2011: number goes from $\bar{z}^{\text{hc}}(0)=3$ and $\bar{z}^{\text{sq}}(0)=4$
2012: to $\bar{z}^{\text{hc}}(1)=2.22$ and $\bar{z}^{\text{sq}}(1)=2.52$;
2013: see Fig.~\ref{FIG:zbar}. The Monte Carlo results are consistent
2014: with our understanding that the quantum fluctuations disrupt the LRO
2015: in inverse proportion to the number of nearest neighbours contributing
2016: to the local staggered mean field at each site.
2017:
2018: %
2019: \begin{figure}
2020: \begin{center}\includegraphics[%
2021: scale=0.85]{fig16-hsd.eps}\end{center}
2022: \caption{\label{FIG:zbar} The disorder-averaged coordination number $\bar{z}(x)$
2023: is plotted as a function of dilution level for infinite square and
2024: honeycomb lattices. The difference between the two lattice types narrows
2025: as $x\rightarrow1$. The undiluted square lattice is 33\% more coordinated
2026: than the honeycomb lattice. At percolation, it is only 12\% more so.}
2027: \end{figure}
2028:
2029:
2030: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2031: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2032: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2033: \section{Summary and concluding remarks}
2034: \label{sec:conclusions}
2035:
2036: In this work we studied the magnetic properties for diluted Heisenberg
2037: models in the honeycomb lattice. Refined results for the
2038: density of states in the square lattice case were also reported.
2039: We have shown that spin wave theory in diluted lattices is quite
2040: successful in describing the magnetic properties of $S>1/2$
2041: systems. On the other hand, for $S=1/2$, spin wave theory breaks down
2042: and one has to approach the problem using a Monte Carlo method.
2043: Contrary to the linear
2044: spin wave method the, the Monte Carlo method does not allow for
2045: the determination of the density of states. Having the advantage
2046: of being rotational invariant by construction, the Monte Carlo
2047: method does not face the problem of the existence of zero energy modes.
2048: We have discussed in detail what is the physics associated with
2049: these modes. In the thermodynamic limit they play the role of
2050: Goldstone modes, trying to restore the rotational symmetry of the problem,
2051: explicitly broken by
2052: the spin wave approximation. We have shown that in a numerical study
2053: these modes can not be included in the calculation of operator
2054: averages, if sensible physical results are to be obtained. This is because
2055: these modes were already used in the construction of the broken symmetry
2056: state, as was first discussed by P.~W.~Anderson in his seminal paper
2057: on spin waves in non-diluted lattices.\cite{AND52}
2058:
2059: Our approach allows us to compute
2060: both the staggered magnetization and the N\'eel temperature as function
2061: of the dilution concentration. In particular, the combination
2062: of spin wave analysis and the recursion method allows for the
2063: calculation of physical quantities virtually in the thermodynamic limit.
2064: This possibility was not used before in similar studies on the square lattice.
2065:
2066: We have used our results to explain the experimental data of two
2067: Heisenberg honeycomb systems: Mn$_{p}$Zn$_{1-p}$PS$_{3}$ (a diluted $S=5/2$
2068: system)
2069: and Ba(Ni$_{p}$Mg$_{1-p}$)$_{2}$V$_{2}$O$_{8}$ (a diluted $S=1$ system).
2070: In the first case, the available experimental and theoretical studies
2071: in the non-diluted regime suggest that second- and third-nearest-neighbor
2072: interactions play a role on the physical properties of the system. This can
2073: be seen from the fact that the measured magnetic moment of the samples
2074: is finite beyond the classical site-dilution percolation threshold.
2075: Our calculation suggests, however, that at low temperatures
2076: and for $p>p_c$ the magnetic moment of these samples can be accounted for on
2077: the basis of a single nearest-neighbor coupling. On the other hand, the
2078: calculation of the N\'eel temperature using a single nearest-neighbor coupling
2079: is underestimated, as it should indeed be case based on the fact that
2080: the magnetic order close to the N\'eel temperature should
2081: have a measurable contribution from the other couplings, which are
2082: not much smaller than the first nearest-neighbor coupling
2083: (the N\'eel temperature for this system using second- and
2084: third-nearest-neighbor interactions will be studied in a future publication).
2085: Simple calculations based on simple (Ising like) mean
2086: filed theories, on the other hand, are very much insensitive, by construction,
2087: to the microscopic details of the system. Therefore, and as long as quantum
2088: fluctuations are not important, a good agreement with the experimental
2089: data should be obtained. This is the case for Mn$_{p}$Zn$_{1-p}$PS$_{3}$,
2090: but not for Ba(Ni$_{p}$Mg$_{1-p}$)$_{2}$V$_{2}$O$_{8}$ since its
2091: much smaller spin brings about the contributions of quantum fluctuations.
2092: In the case of the system Ba(Ni$_{p}$Mg$_{1-p}$)$_{2}$V$_{2}$O$_{8}$,
2093: there are, unfortunately, no measurement of its magnetic moment in the
2094: diluted phase, however, the N\'eel temperature as function of dilution
2095: is known from thermodynamic measurements. Our results show
2096: that in this case, most likely, only the first-nearest-neighbor coupling
2097: (and a very small magnetic anisotropy) are needed to describe the behavior
2098: of the N\'eel temperature upon dilution. It would be important if further
2099: investigations on this system could be performed in the future.
2100:
2101:
2102: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2103: \subsection*{Acknowledgements}
2104:
2105: Some of the understanding presented in this paper
2106: on the physics of the zero modes reflects a number of
2107: enlightening discussions with J. B. M. Lopes dos Santos,
2108: for which the authors are grateful.
2109: We thank A. H. Castro Neto for illuminating conversations
2110: of the physics of the 2D antiferromagnet in a square lattice. E.V.C.
2111: acknowledges the Quantum Condensed Matter Theory Group at Boston University,
2112: Boston, MA, U.S.A., for the hospitality, and the financial support of
2113: Funda\c{c}\~ao para a Ci\^encia e a Tecnologia through Grant
2114: Ref. SFRH/BD/13182/2003. N.M.R.P. is thankful to the Quantum Condensed
2115: Matter visitors program at Boston University, Boston, MA, U.S.A., to the
2116: visitors program at the Max-Planck-Institut f\"ur Physik komplexer Systeme,
2117: Dresden, Germany, and to Funda\c{c}\~ao para a Ci\^encia e a Tecnologia for
2118: a sabbatical grant. E.V.C., N.M.R.P. and J.L.B.L.S. were additionally financed
2119: by FCT and EU through POCTI (QCAIII).
2120:
2121:
2122:
2123: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2124: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2125: \appendix
2126:
2127: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2128: \section{Diagonalization of $H_{\mathbf{k}=\mathbf{0}}$}
2129: \label{sec:App-diagHk0}
2130:
2131: The $\mathbf{k}=\mathbf{0}$ term in Hamiltonian \eqref{eq:Hk} can
2132: be expressed as
2133: \begin{equation}
2134: H_{\mathbf{0}}=JSz\left(h_{a}(a_{\mathbf{0}}a_{\mathbf{0}}^{\dagger}
2135: +b_{\mathbf{0}}^{\dagger}b_{\mathbf{0}})+a_{\mathbf{0}}b_{\mathbf{0}}
2136: +b_{\mathbf{0}}^{\dagger}a_{\mathbf{0}}^{\dagger}\right).
2137: \label{eq:Hk0}
2138: \end{equation}
2139: This is a standard bilinear model with two coupled modes, which is
2140: straightforwardly diagonalized through a Bogoliubov-Valatin transformation
2141: (Eq.~\eqref{eq:BVT}) when $h_{a}>1$. In the isotropic $h_{a}=1$
2142: case it has an infinite number of eigenstates with a continuum energy
2143: spectrum.
2144:
2145: Let us define the following canonical transformation,
2146: \begin{align}
2147: a_{\mathbf{0}} & = \hat{q}_{1}+i\hat{p}_{1}\label{eq:a0q1p1}\,,\\
2148: b_{\mathbf{0}} & = \hat{q}_{2}+i\hat{p}_{2}\,.\label{eq:b0q2p2}
2149: \end{align}
2150: We use the \emph{hat} notation to distinguish the operators from
2151: their eigenvalues. The new generalized {}``position'' $\hat{q}$
2152: and {}``momentum'' $\hat{p}$ operators satisfy the usual commutation
2153: relations:
2154: \begin{alignat}{3}
2155: \bigl[a_{\mathbf{0}},a_{\mathbf{0}}^{\dagger}\bigr]&=1 &
2156: \quad &\Longrightarrow \quad &
2157: \bigl[\hat{q}_{1},\hat{p}_{1}\bigr]&=\frac{i}{2}\,;\label{eq:q1p1cr}\\
2158: \bigl[b_{\mathbf{0}},b_{\mathbf{0}}^{\dagger}\bigr]&=1 &
2159: \quad &\Longrightarrow \quad &
2160: \bigl[\hat{q}_{2},\hat{p}_{2}\bigr]&=\frac{i}{2}\,.\label{eq:q2p2cr}
2161: \end{alignat}
2162: After simple algebra we find that the $h_{a}=1$ Hamiltonian \eqref{eq:Hk0}
2163: can be written in terms of the new operators $\hat{q}$'s and $\hat{p}$'s as
2164: \begin{equation}
2165: H_{\mathbf{0}}=JSz\left[(\hat{q}_{1}+\hat{q}_{2})^{2}+(\hat{p}_{1}
2166: -\hat{p}_{2})^{2}\right].
2167: \label{eq:H0qp}
2168: \end{equation}
2169: The variables $\hat{q}_{1}+\hat{q}_{2}$ and $\hat{p}_{1}-\hat{p}_{2}$
2170: can be interpreted as the center of mass position and the relative
2171: momentum, respectively, of a two particle system, therefore commuting
2172: with each other
2173: \begin{equation}
2174: [\hat{q}_{1}+\hat{q}_{2},\,\hat{p}_{1}-\hat{p}_{2}]=\frac{i}{2}-\frac{i}{2}=0.
2175: \label{eq:QPcr}
2176: \end{equation}
2177: Thus the eigenfunctions of Hamiltonian \eqref{eq:Hk0} are given
2178: in as products of the eigenstates of the operator $\hat{q}_{1}+\hat{q}_{2}$
2179: with eigenstates of the operator $\hat{p}_{1}-\hat{p}_{2}$,
2180: \begin{equation}
2181: \Psi_{Q,P}(q_{1},q_{2})=\delta(q_{1}+q_{2}-Q)e^{i\frac{P}{2}(q_{1}-q_{2})},
2182: \label{eq:eigfuncHk0}
2183: \end{equation}
2184: and the aforementioned continuum spectrum is given by
2185: \begin{equation}
2186: E_{Q,P}=JSz(Q^{2}+P^{2}).
2187: \label{eq:Eqp}
2188: \end{equation}
2189:
2190:
2191: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2192: \section{Anderson broken symmetry analysis}
2193: \label{sec:App-Anderson}
2194:
2195: In this appendix we closely follow the ideas developed by
2196: P.~W.~Anderson\cite{AND52} to show that the ground state of an
2197: antiferromagnet should display broken spin rotational symmetry,
2198: even in the absence of any anisotropy.
2199:
2200: As was shown in Appendix \ref{sec:App-diagHk0} the operators
2201: $\hat{q}_{1}+\hat{q}_{2}$ and $\hat{p}_{1}-\hat{p}_{2}$ are constants of
2202: the motion in a system described by the isotropic Hamiltonian \eqref{eq:Hk0},
2203: having well defined expectation values with zero dispersion. From definitions
2204: \eqref{eq:a0q1p1} and \eqref{eq:b0q2p2}, and Eqs.~\eqref{eq:akbk}
2205: and \eqref{eq:Sapprox}, it can be easily seen that
2206: \begin{align}
2207: \hat{q}_{1}+\hat{q}_{2} & = \frac{1}{\sqrt{2SN_{a}}}S_{\text{tot}}^{x}\,,
2208: \label{eq:QSx}\\
2209: \hat{p}_{1}-\hat{p}_{2} & = \frac{1}{\sqrt{2SN_{a}}}S_{\text{tot}}^{y}\,,
2210: \label{eq:PSy}
2211: \end{align}
2212: which explains the constant of motion character of the
2213: $\hat{q}_{1}+\hat{q}_{2}$ and $\hat{p}_{1}-\hat{p}_{2}$ operators
2214: ($\mathbf{S}_{\text{tot}}$ is a constant of motion of the original
2215: Heisenberg model). The uncertainty relation ensures us that their
2216: canonical conjugates counterparts will have divergent dispersion.
2217: As for Eq.~\eqref{eq:QSx} and \eqref{eq:PSy} it is not difficult to
2218: show that the canonical conjugates of $\hat{p}_{1}-\hat{p}_{2}$
2219: and $\hat{q}_{1}+\hat{q}_{2}$ are, respectively,
2220: \begin{align}
2221: \hat{q}_{1}-\hat{q}_{2} & = \frac{1}{\sqrt{2SN_{a}}}\biggl(\sum_{i\in A}
2222: S_{i}^{a,x}-\sum_{i\in B}S_{i}^{b,x}\biggr), \label{eq:q-mx}\\
2223: \hat{p}_{1}+\hat{p}_{2} & = \frac{1}{\sqrt{2SN_{a}}}\biggl(\sum_{i\in A}
2224: S_{i}^{a,y}-\sum_{i\in B}S_{i}^{b,y}\biggr). \label{eq:p-my}
2225: \end{align}
2226:
2227: We want to know how much energy is needed to form a wave packet with
2228: states \eqref{eq:eigfuncHk0} (above the ground state), such as the
2229: expectation values of operators $\hat{q}_{1}-\hat{q}_{2}$ and
2230: $\hat{p}_{1}+\hat{p}_{2}$ have finite dispersion. Let us limit the
2231: fluctuations of the expectation value $\left\langle
2232: \hat{q}_{1}-\hat{q}_{2}\right\rangle $ to the
2233: range $\Delta_{\hat{q}_{1}-\hat{q}_{2}}$. From the uncertainty relation
2234: the expectation value of $\hat{p}_{1}-\hat{p}_{2}$ must now have
2235: a nonzero dispersion, whose magnitude is given by
2236: \begin{equation}
2237: \Delta_{\hat{p}_{1}-\hat{p}_{2}}
2238: \approx\frac{1}{2\Delta_{\hat{q}_{1}-\hat{q}_{2}}}\,.
2239: \label{eq:incert}
2240: \end{equation}
2241: Thus, to limit $\left\langle \hat{q}_{1}-\hat{q}_{2}\right\rangle $
2242: to the range $\Delta_{\hat{q}_{1}-\hat{q}_{2}}$ we need
2243: \begin{equation}
2244: E_{lim}\simeq\frac{JSz}{4\Delta_{\hat{q}_{1}-\hat{q}_{2}}^{2}}\,,
2245: \label{eq:Elim-mx}
2246: \end{equation}
2247: relatively to the ground state energy. Analogously, to limit
2248: $\left\langle \hat{p}_{1}+\hat{p}_{2}\right\rangle $ to the range
2249: $\Delta_{\hat{p}_{1}+\hat{p}_{2}}$ we need
2250: \begin{equation}
2251: E_{lim}\simeq\frac{JSz}{4\Delta_{\hat{p}_{1}+\hat{p}_{2}}^{2}}\,.
2252: \label{eq:Elim-my}
2253: \end{equation}
2254: Defining the the mean square amplitudes of the quantities \eqref{eq:magx}
2255: and \eqref{eq:magy},
2256: \begin{align}
2257: \sigma_{x}^{2} & = \frac{1}{(2N_{a})^{2}}\left\langle \left(\sum_{i\in A}
2258: S_{i}^{a,x}-\sum_{i\in B}S_{i}^{b,x}\right)^{\!\!2\,}\right\rangle\,,
2259: \label{eq:VARmagx}\\
2260: \sigma_{y}^{2} & = \frac{1}{(2N_{a})^{2}}\left\langle \left(\sum_{i\in A}
2261: S_{i}^{a,y}-\sum_{i\in B}S_{i}^{b,y}\right)^{\!\!2\,}\right\rangle\,,
2262: \label{eq:VARmagy}
2263: \end{align}
2264: we find from \eqref{eq:q-mx} and \eqref{eq:p-my} that
2265: \begin{align}
2266: \Delta_{\hat{q}_{1}-\hat{q}_{2}}^{2} & = \frac{2N_{a}}{S}\sigma_{x}^{2}\,,
2267: \label{eq:VARq1-q2}\\
2268: \Delta_{\hat{p}_{1}+\hat{p}_{2}}^{2} & = \frac{2N_{a}}{S}\sigma_{y}^{2}\,,
2269: \label{eq:VARp1+p2}
2270: \end{align}
2271: (note that $\hat{q}_{1}-\hat{q}_{2}$ and $\hat{p}_{1}+\hat{p}_{2}$
2272: have zero expectation value). Inserting \eqref{eq:VARq1-q2} and
2273: \eqref{eq:VARp1+p2} in Eqs.~\eqref{eq:Elim-mx} and \eqref{eq:Elim-my}
2274: it can be seen that to limit $\sigma_{x}$ or $\sigma_{y}$ to a finite value we
2275: only need an excess energy of magnitude $1/N_{a}$, and hence negligible
2276: in the thermodynamic limit. As pointed out by Anderson, we can even
2277: limit $\sigma_{x}$ or $\sigma_{y}$ to values of magnitude
2278: $1/N_{a}^{\frac{1}{2}+\alpha}$, with $\alpha>0$, requiring no energy when
2279: $N_{a}\rightarrow\infty$.
2280:
2281:
2282: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2283: \bibliographystyle{apsrev}
2284: \bibliography{./Bibliography/computationalphys,./Bibliography/condmattbooks,./Bibliography/magsitedilutexp,./Bibliography/magsitediluttheory,./Bibliography/recursionmethod}
2285:
2286: \end{document}
2287: