math0008243/cep.tex
1: \documentclass{amsart}
2: 
3: % This is the LaTeX source file for ``Local statistics for random
4: % domino tilings of the Aztec diamond'' by Henry Cohn, Noam Elkies,
5: % and James Propp.  To process it correctly, you must have 7 files
6: % containing the postscript figures.  These files are called
7: % tiling.ps, aztec2.ps, heights.ps, flat.ps, aug32.ps, daug32.ps, and
8: % herringbone.ps.  In order to produce postscript output containing
9: % the figures, these 7 files must be in the directory in which the
10: % LaTeX processing is being done.  If any of the file names are
11: % changed (or the files are in another directory), then the \PSbox
12: % commands must be changed to reflect the new file names.  For
13: % example, if the file aug32.ps is instead called Figures/aug.ps, then
14: % the line that currently reads
15: % \PSbox{aug32.ps}{250pt}{260pt}
16: % should be changed to
17: % \PSbox{Figures/aug.ps}{250pt}{260pt}
18: 
19: % This is the final version (including typo corrections to page proofs).
20: 
21: \newcommand{\PSbox}[3]{\mbox{\rule{0in}{#3}\special{psfile=#1}\hspace{#2}}}
22: % For including postscript figures.
23: 
24: \theoremstyle{plain}
25: \newtheorem{thm}{Theorem}
26: \newtheorem{cor}[thm]{Corollary}
27: \newtheorem{lem}[thm]{Lemma}
28: \newtheorem{prop}[thm]{Proposition}
29: 
30: \numberwithin{equation}{section}
31: 
32: \newcommand{\thmref}[1]{Theorem~\ref{#1}}
33: \newcommand{\propref}[1]{Proposition~\ref{#1}}
34: \newcommand{\corref}[1]{Corollary~\ref{#1}}
35: \newcommand{\secref}[1]{\S\ref{#1}}
36: \newcommand{\lemref}[1]{Lemma~\ref{#1}}
37: 
38: 
39: \newcommand{\Z}    {{\mathbb Z}}
40: \newcommand{\C}    {{\mathbb C}}
41: 
42: \newcommand{\Cr}   {{\textup{Cr}}}
43: \newcommand{\Pl}   {{\textup{P}}}
44: \newcommand{\Plt}  {{\textup{\~P}}}
45: 
46: \newcommand{\Exp}  {{\textup{Exp}}}
47: \newcommand{\Var}  {{\textup{Var}}}
48: \newcommand{\Prob} {{\textup{Prob}}}
49: 
50: \title[Local statistics for random domino tilings]{Local statistics
51: for random domino tilings of the Aztec diamond} 
52: \author{Henry Cohn}
53: \address{Department of Mathematics, Harvard University}
54: \email{cohn@math.harvard.edu} 
55: \author{Noam Elkies} 
56: \address{Department of Mathematics, Harvard University} 
57: \email{elkies@math.harvard.edu}
58: \author{James Propp} 
59: \address{Department of Mathematics, MIT}
60: \email{propp@math.mit.edu} 
61: \thanks{The second author was supported in part by an NSF Presidential
62: Young Investigator Award and by a Fellowship from the Packard
63: Foundation. The third author was supported in part by NSA grant
64: MDA904-92-H-3060 and NSF grant DMS 9206374, and by a career
65: development grant from the M.I.T.\ Class of 1922.} 
66: \subjclass{Primary 60K35, 82B20; Secondary 05A16, 60C05}
67: \date{April, 1996}
68: \begin{document}
69: 
70: \begin{abstract}
71: We prove an asymptotic formula for the probability that, if one
72: chooses a domino tiling of a large Aztec diamond at random according
73: to the uniform distribution on such tilings, the tiling will contain a
74: domino covering a given pair of adjacent lattice squares.  This
75: formula quantifies the effect of the diamond's boundary conditions on
76: the behavior of typical tilings; in addition, it yields a new proof of
77: the arctic circle theorem of Jockusch, Propp, and Shor.  Our approach
78: is to use the saddle point method to estimate certain weighted sums of
79: squares of Krawtchouk polynomials (whose relevance to domino tilings
80: is demonstrated elsewhere), and to combine these estimates with some
81: exponential sum bounds to deduce our final result.  This approach
82: generalizes straightforwardly to the case in which the probability
83: distribution on the set of tilings incorporates bias favoring
84: horizontal over vertical tiles or vice versa.  We also prove a fairly
85: general large deviation estimate for domino tilings of
86: simply-connected planar regions that implies that some of our results
87: on Aztec diamonds apply to many other similar regions as well.
88: \end{abstract}
89: 
90: \maketitle
91: 
92: \section{Introduction}
93: \label{sec-introduction}
94: 
95: \subsection{Statement of the main theorem.}
96: \label{ssec-statement}
97: 
98: Random domino tilings of finite regions often exhibit surprising
99: statistical heterogeneity.  Such heterogeneity would be expected in
100: the vicinity of the boundary, but in fact the presence of a boundary
101: can make its influence felt well into the interior of the region.  The
102: research that led to this article is part of an ongoing effort to
103: understand this phenomenon.  The results proved here are the first to
104: give a precise description of how local statistics for domino tilings
105: can vary continuously throughout a region in response to the
106: imposition of specific boundary conditions.
107: 
108: Those who study random tilings of finite regions (in the plane) by
109: dominos have tended to focus on regions that are rectangles of even
110: area.  In particular, Burton and Pemantle \cite{local} have done an
111: intensive analysis of the small-scale structure of such tilings.
112: Their work shows that once one gets away from the boundary of the
113: rectangle, random tilings tend to exhibit statistical isotropy.  Among
114: all random processes that take their values in the set of domino
115: tilings of the plane, the Burton-Pemantle process has maximal entropy,
116: and it is unique in this regard; for this reason alone, it is worth
117: further study.
118: 
119: However, if one looks at random domino tilings of tileable finite
120: regions in general, one finds that local behavior far from the
121: boundary need not be governed by maximal entropy statistics, but can
122: look very different.  Moreover, the local behavior seen in one part of
123: the region is in general different from local behaviors seen
124: elsewhere.
125: 
126: One especially tractable proving ground for the study of this
127: statistical heterogeneity has been the family of finite regions known
128: as Aztec diamonds, introduced and studied in \cite{alternating}.
129: Figure~\ref{fig-aztec} shows an Aztec diamond of order 64 tiled
130: randomly by dominos.  In general, the Aztec diamond of order $n$ can
131: be defined as the union of those lattice squares whose interiors lie
132: inside the region $\{(x,y): x+y \leq n+1\}$.
133: 
134: \begin{figure}
135: \begin{center}
136: \PSbox{tiling.ps hoffset=0 voffset=0}{350pt}{350pt}
137: \end{center}
138: \caption{A random domino tiling of an Aztec diamond of order 64.}
139: \label{fig-aztec}
140: \end{figure}
141: 
142: It was shown in \cite{circle} (and will be proved in
143: subsection~\ref{ssec-ACT}~by different methods) that, asymptotically,
144: the circle inscribed in the Aztec diamond of order $n$ serves as a
145: boundary between domains of qualitatively different behavior.  We call
146: this circle the arctic circle, because, as one can see from
147: Figure~\ref{fig-aztec}, the dominos outside the arctic circle are
148: frozen into a brickwork pattern.  To state the theorem more precisely,
149: we impose a checkerboard coloring on the Aztec diamond of order $n$,
150: so that the leftmost square in each row in the top half of the diamond
151: is white.  We say a horizontal domino is {\sl north-going} or {\sl
152: south-going} according to whether its leftmost square is white or
153: black, and we say a vertical domino is {\sl west-going} or {\sl
154: east-going} according to whether its upper square is white or black.
155: (The motivation for this terminology comes from the ``domino
156: shuffling'' algorithm introduced in \cite{alternating} and used in
157: both \cite{circle} and \cite{gip}; this algorithm permits one to
158: generate random domino tilings of Aztec diamonds in such a way that
159: every possible tiling has the same probability of arising as every
160: other, and indeed it was this algorithm that we used to generate the
161: tiling shown in Figure~\ref{fig-aztec}.)
162: 
163: Say that two dominos are adjacent if they share an edge (i.e., their
164: boundaries overlap on a segment of length $1$ or more), and say that a
165: domino is adjacent to the boundary of the Aztec diamond if it shares
166: an edge with the boundary.  We define the {\sl north polar region} as
167: the union of those north-going dominos that are each connected to the
168: boundary by a sequence of adjacent north-going dominos.  The south,
169: west, and east polar regions are defined similarly, and the {\sl
170: temperate zone} is the union of those dominos that belong to none of
171: the four polar regions.
172: 
173: The arctic circle theorem of \cite{circle} states that for every
174: $\varepsilon > 0$, if one takes $n$ sufficiently large, then for all
175: but an $\varepsilon$ fraction of the domino tilings of the diamond of
176: order $n$, the border of the temperate zone stays within distance
177: $\varepsilon n$ of the circle of radius $n/\sqrt{2}$ with center
178: $(0,0)$.  In particular, this implies that if one increases the radius
179: of the circle by $\varepsilon n$, then with probability greater than
180: $1-\varepsilon$, in each of the four regions in the Aztec diamond that
181: lie outside the enlarged disk, all dominos are aligned with their
182: neighbors in brickwork patterns.  The theorem also implies that if one
183: decreases the radius of the disk by $\varepsilon n$, then within the
184: shrunken disk dominos with different orientations are in some sense
185: interspersed among one another (with probability greater than
186: $1-\varepsilon$); however, the theorem by itself gives no information
187: on their distribution.
188: 
189: In \thmref{main} of this article we will give a quantitative analysis
190: of the behavior of random tilings in the inner, disorderly zone.  In
191: particular, we will give an asymptotic formula for the proportion of
192: domino tilings of the Aztec diamond of order $n$ that contain a domino
193: at a specified location, i.e., the {\sl placement probability} for
194: that location.  This formula depends only on the orientation of the
195: domino, its parity relative to the natural checkerboard coloring of
196: the Aztec diamond, and the relative position of the domino within the
197: Aztec diamond (in normalized coordinates).  One consequence of our
198: formula is that random domino tilings exhibit ``total statistical
199: heterogeneity'' within the central zone.  That is to say, any two
200: patches within the temperate zone that are macroscopically separated
201: (i.e., separated by a distance on the order of $n$) will exhibit
202: distinct statistics.  (For a precise statement, see
203: subsection~\ref{ssec-heterogeneity}.)
204: 
205: Our work builds on the generating functions derived in \cite{gip}.
206: One of them is a rational function in three variables whose
207: coefficients are the placement probabilities for which an asymptotic
208: formula is sought.  The authors of the earlier article carried out a
209: relatively straightforward complex integration to calculate
210: coefficients corresponding to dominos in the $2\times2$ block in the
211: middle of the Aztec diamond; the resulting exact formula implies that
212: in a diamond of order $n$, these placement probabilities are $\frac14
213: + O(\frac1n)$.  In the present article we will apply the saddle point
214: method to estimate contour integrals associated with more general
215: coefficients of a related generating function (also derived in
216: \cite{gip}).
217: 
218: We can now prepare to state our main result.  We call the union of two
219: adjacent squares in the Aztec diamond a {\sl domino space\/}, to avoid
220: confusion between actual dominos occurring in a particular tiling and
221: the locations in which dominos can occur.  Domino spaces are
222: classified as north-going, south-going, west-going, or east-going in
223: the obvious way, so that for instance a domino is north-going if and
224: only if it occupies a north-going domino space.  Because of symmetry,
225: we lose no generality by focusing on the placement probabilities
226: associated with north-going domino spaces.  The midpoint of the bottom
227: edge of each north-going domino space is some point $(\ell,m)$ with
228: $|\ell|+|m| \leq n-1$.  We call this the {\sl location} of the
229: north-going domino space.  Normalizing by dividing by $n$, we obtain
230: some point $(x,y)$ with $|x|+|y| < 1$.  We call this the {\sl
231: normalized location} of the north-going domino space.
232: 
233: \begin{thm}
234: \label{main}
235: Let $U$ be an open set containing the points $(\pm\frac12, \frac12)$.
236: If $(x,y)$ is the normalized location of a north-going domino space in
237: the Aztec diamond of order $n$, and $(x,y) \not\in U$, then, as $n
238: \rightarrow \infty$, the placement probability at $(x,y)$ is within
239: $o(1)$ of ${\mathcal P}(x,y)$, where
240: $$
241: {\mathcal P}(x,y) = 
242: \begin{cases}
243: 0&\hbox{if $x^2+y^2\ge\frac12$ and $y<\frac12$,}\\
244: 1&\hbox{if $x^2+y^2\ge\frac12$ and $y>\frac12$, and}\\
245: \frac{1}{2}+\frac{1}{\pi}\tan^{-1}
246: \left(\frac{2y-1}{\sqrt{1-2x^2-2y^2}}\right)
247: &\hbox{if $x^2+y^2<\frac12$}.\\
248: \end{cases}
249: $$
250: The $o(1)$ error bound is uniform in $(x,y)$ (for $(x,y) \not\in U$).
251: \end{thm}
252: 
253: Similarly, the south-going, east-going, and west-going placement
254: probabilities near $(x,y)$ are approximated by ${\mathcal P}(-x,-y)$,
255: ${\mathcal P}(-y,x)$, and ${\mathcal P}(y,-x)$, respectively.  This
256: follows from \thmref{main} by rotational symmetry.
257: 
258: The organization of the rest of this article is as follows.  
259: 
260: In the remainder of Section~\ref{sec-introduction}, we discuss some
261: qualitative features of the main theorem and give some preliminaries
262: for the proof.  In Section~\ref{sec-creation}, we use the saddle point
263: method to derive asymptotic estimates for certain numbers known as
264: creation rates, which give placement probabilities when summed
265: appropriately.  In Section~\ref{sec-placement}, we use this result to
266: estimate, modulo an error term, the north-going placement
267: probabilities.  In Section~\ref{sec-exponential}, we use techniques
268: {}from the theory of exponential sums to justify our bound for the error
269: term.  This completes the proof of the theorem away from the boundary
270: of the diamond; Section~\ref{sec-conclusion}~provides the final
271: arguments that handle locations near the boundary.
272: 
273: Section~\ref{sec-consequences}~discusses some consequences of the
274: theorem.  In particular, by taking a detour through the theory of
275: domino tilings in general, we show that some consequences of the
276: arctangent formula apply not only to the particular shape we call the
277: Aztec diamond but also to slightly deformed versions of this shape
278: (\propref{robustness}).  We also give a new proof of the arctic circle
279: theorem and a large deviation estimate for certain properties of
280: random tilings of simply-connected finite regions
281: (\thmref{variancethm} and \propref{large}). 
282: Section~\ref{sec-further}~briefly sketches how the method of proof of
283: Theorem~\ref{main} can be adapted to handle the more general case of
284: random domino tilings when there is a bias in favor of one domino
285: orientation over the other (horizontal versus vertical).  We conclude
286: in Section~\ref{sec-speculations}~with speculations and open
287: questions.
288: 
289: For a treatment of the probabilistic preliminaries needed for
290: Section~\ref{sec-consequences}, see \cite{durrett}.
291: 
292: \subsection{Features of the result.}
293: \label{ssec-features}
294: 
295: As a first comment on the qualitative features of this formula, we
296: point out the continuity of the formula for ${\mathcal P}(x,y)$
297: (except at $(\pm\frac12,\frac12)$).  Indeed, if we had been so naive
298: as to ask for an asymptotic formula for the placement probabilities
299: for {\sl all} horizontal domino spaces in an asymptotically small
300: patch of the Aztec diamond (south-going as well as north-going), we
301: would not get a single value at all but rather a pair of values,
302: namely ${\mathcal P}(x,y)$ and ${\mathcal P}(-x,-y)$, which are not in
303: general equal.  That is, the local statistics are not even
304: approximately invariant under translations that exchange the two
305: color-classes.  It is therefore all the more pleasant that the local
306: statistics are asymptotically invariant under translations that
307: preserve the two color-classes (at least, they are invariant if, in
308: discussing local statistics, we confine ourselves to placement
309: probabilities, and do not inquire about correlations between
310: placements).
311: 
312: Another important feature of the formula is the singular behavior that
313: occurs near the normalized locations $(\pm \frac12, \frac12)$, which
314: we can explain as follows.  In \cite{alternating} it is shown that the
315: Aztec diamond of order $n$ has exactly $2^{n(n+1)/2}$ domino tilings,
316: and a formula derived in that article (formula~(7) of Section~4) can
317: be used to show that for $0 \leq k \leq n$, exactly $\binom{n}{k}
318: 2^{n(n-1)/2}$ of the tilings have horizontal dominos covering the
319: leftmost squares in the first $k$ rows from the top and have vertical
320: dominos covering the leftmost squares in the next $n-k$ rows.  Thus,
321: the placement probability associated with the leftmost north-going
322: domino space in the $k$th row is exactly the sum
323: \[ 2^{-n} \sum_{i=k}^n \binom{n}{i}.\]
324: This sum is very close to 1 for $k -\frac{n}{2} \ll - \sqrt{n}$ and
325: very close to 0 for $k -\frac{n}{2} \gg \sqrt{n}$; macroscopically
326: speaking, the placement probability jumps from 1 to 0 discontinuously.
327: It might be possible to analyze the limiting behavior of the placement
328: probabilities in the vicinity of the singularities under suitable
329: scaling, but we do not explore this possibility here.
330: 
331: An easily-understood symmetry property of ${\mathcal P}(\cdot,\cdot)$
332: is the fact that
333: \begin{equation} \label{reflection}
334: {\mathcal P}(x,y) = {\mathcal P}(-x,y).  
335: \end{equation}
336: This is a consequence of the fact that reflecting a domino tiling
337: through the line $x=0$ carries north-going domino spaces to
338: north-going domino spaces.  A further identity satisfied by ${\mathcal
339: P}(\cdot,\cdot)$ is the relation
340: \begin{equation} 
341: \label{rotation}
342: {\mathcal P}(x,y) + {\mathcal P}(-y,x) + {\mathcal P}(-x,-y) +
343: {\mathcal P}(y,-x) = 1.
344: \end{equation}
345: To see why this is true, one need only observe that the four domino
346: spaces that contain a particular lattice square (fewer, if the square
347: is on the boundary) must have placement probabilities that sum to 1.
348: 
349: A subtler consequence of Theorem~\ref{main} is the fact that the level
350: sets of ${\mathcal P}(x,y)$ (for probabilities strictly between 0 and
351: 1) are arcs of ellipses.  More specifically, for $0 < p < 1$ the level
352: set $\{(x,y): {\mathcal P}(x,y)=p\}$ and the level set $\{(x,y):
353: {\mathcal P}(x,y) =1-p\}$, together with the singular points $(\pm
354: \frac12, \frac12)$, form an ellipse tangent to the boundary of the
355: diamond at the two singular points.  As $p \rightarrow 0$ (or $p
356: \rightarrow 1$), the ellipse becomes the inscribed circle, which is
357: the zero-set of the function $2x^2+2y^2-1$; in the case $p=\frac12$,
358: the ellipse degenerates into the line segment joining the two singular
359: points, which is the part of the zero-set of the function $(2y-1)^2$
360: lying inside the Aztec diamond; and in general, the ellipse will be
361: the zero-set of some convex combination of $2x^2+2y^2-1$ and
362: $(2y-1)^2$.  The point $(0,0)$ lies on the level set $p=\frac14$,
363: which is an arc of an ellipse; the complementary arc of the ellipse is
364: the level set $p=\frac34$, and the point on this arc opposite $(0,0)$
365: is the point $(0,\frac23)$.  The situation is depicted schematically
366: in Figure~\ref{fig-ellipses}.
367: 
368: \begin{figure}
369: \begin{center}
370: \PSbox{aztec2.ps hoffset=-25 voffset=-20}{3.8in}{3.8in}
371: \end{center}
372: \caption{Level curves of north-going placement probabilities.}
373: \label{fig-ellipses}
374: \end{figure}
375: 
376: \subsection{Preparation for the proof.}
377: \label{ssec-preparation}
378: 
379: Recall that, under the original (unnormalized) coordinate system, each
380: north-going domino space in an Aztec diamond of order $n$ is assigned
381: some location $(\ell,m)$ with $|\ell|+|m| \leq n-1$.  It is easy to
382: check that $\ell+m$ must have the same parity as $n-1$.  Define
383: $\Pl(\ell,m;n)$ as the probability that a random domino tiling of the
384: Aztec diamond of order $n$ will have a domino occupying the
385: north-going domino space at location $(\ell,m)$; for $|\ell|+|m| >
386: n-1$, or $\ell+m \not \equiv n-1 \pmod{2}$, define $\Pl(\ell,m;n)=0$.
387: For instance, we have $\Pl(0,0;1)=\frac12$, $\Pl(0,1;2)=\frac34$, and
388: $\Pl(0,-1;2)=\Pl(1,0;2)=\Pl(-1,0;2)=\frac14$.
389: 
390: Define 
391: \begin{equation}
392: \label{creation}
393: \Cr(\ell,m;n) = 2(\Pl(\ell,m;n) - \Pl(\ell,m-1;n-1)).
394: \end{equation}
395: This quantity is called the {\sl net creation rate} at location
396: $(\ell,m)$, but the reason for this name and the interpretation of the
397: quantity in terms of domino shuffling are not needed for our purposes.
398: (For the motivation, see \cite{gip}.)
399: 
400: Define $c(a,b;n)$ to be the coefficient of $z^a$ in
401: $(1+z)^{n-b}(1-z)^b$.  (Note that $c(a,b;n)$ is the Krawtchouk
402: polynomial $P_a$ evaluated at $b$.  For information about Krawtchouk
403: polynomials, see \cite[p.~130]{codes}.)  Our proof of \thmref{main}
404: will be based on the following result from \cite{gip}:
405: 
406: \begin{prop}
407: \label{gessel}
408: Let $n>0$.  Suppose $\ell$ and $m$ are integers with $\ell+m \equiv n
409: \pmod{2}$ and $|\ell|+|m| \le n$.  If we let $a=(\ell+m+n)/2$ and
410: $b=(\ell-m+n)/2$, then
411: $$
412: \Cr(\ell,m;n+1)=c(a,b;n)c(b,a;n)/2^n.
413: $$
414: For other integers $\ell$ and $m$, we have $\Cr(\ell,m;n+1)=0$.
415: \end{prop}
416: 
417: This proposition implies that the creation rates are non-negative, if
418: we use the identity $c(b,a;n)b!(n-b)!=c(a,b;n)a!(n-a)!$.  (This
419: identity is a standard fact about Krawtchouk polynomials, and follows
420: immediately from Theorem~17 on page~152 of \cite{codes}.)  When
421: combined with \propref{gessel}, the identity implies that
422: $\Cr(\ell,m;n+1)$ is a positive factor times the square of a
423: Krawtchouk polynomial, and hence that $\Cr(\ell,m;n+1) \ge 0$.  Note
424: that this inequality, together with (\ref{creation}), yields
425: \begin{equation}
426: \label{incbound}
427: \Pl(\ell,m;n) \le \Pl(\ell,m+h;n+h)
428: \end{equation}
429: for $h \ge 0$ by induction on $h$.
430: 
431: We will also need the following result on exponential sums.  
432: 
433: \begin{thm}[Kusmin-Landau]
434: \label{kusmin-landau}
435: Let $||\cdot||$ denote the distance to the nearest integer, $I$ be an
436: interval, and $f$ be a real-valued function on $I$.  If $f$ is
437: continuously differentiable, $f'$ is monotonic, and $ ||f'|| \ge
438: \lambda > 0$ on $I$, then
439: $$\sum_{n \in I\cap\Z}\exp(2\pi i \,f(n)) = O(\lambda^{-1}).$$
440: The constant implicit in the
441: $O(\lambda^{-1})$ term does not depend on $I$.
442: \end{thm}
443: 
444: \noindent A proof can be found in \cite[p.~7]{expsums}.
445: 
446: \subsection{Outline of the Proof of Theorem~1}
447: \label{ssec-outline}
448: 
449: We begin the proof of \thmref{main} by using (\ref{creation}) to write
450: the placement probabilities as sums of creation rates, which gives the
451: formula
452: \begin{equation}
453: \label{crsum}
454: \Pl(\ell,m;n) = \frac{1}{2}\sum_{k \ge 0}{\Cr(\ell,m-k;n-k)}.
455: \end{equation}
456: (Note that the remark after \propref{gessel} shows that, as claimed in
457: the abstract, this is a weighted sum of squares of Krawtchouk
458: polynomials.)  We will estimate the creation rates, and then use our
459: estimate to prove the asymptotic formula for the placement
460: probabilities.
461: 
462: Because of \propref{gessel}, to estimate the creation rates it
463: suffices to approximate the coefficients of the polynomials
464: $(1+z)^{n-b}(1-z)^b$.  To do this, we write the coefficients as
465: contour integrals in the usual way, and then apply the saddle point
466: method to these integrals.  Sufficiently far outside the arctic
467: circle, this method shows that the creation rates are exponentially
468: small in $n$ (\propref{expsmall}); sufficiently far inside, it
469: approximates them by a well-behaved function times an oscillating
470: factor (\propref{crest}).  With additional work, it might be possible
471: to obtain a uniform estimate over the entire Aztec diamond, but we can
472: make do with just these estimates.
473: 
474: We would then like to substitute our creation rate estimates into
475: (\ref{crsum}) and convert the sum to an integral to determine its
476: asymptotics.  If we are willing to be unrigorous, we can wishfully
477: replace the oscillatory cosine-squared factor in \propref{crest} by
478: its mean value $\frac12$, obtaining (for locations inside the
479: inscribed circle)
480: \begin{eqnarray*}
481: \Pl(\ell, m; n)  
482: & \approx & \frac12 \sum_{k = 0}^{t_{{\rm max}}} 
483: 	\frac{2}{\pi \sqrt{(n-k)^2-2\ell^2-2(m-k)^2}} \\
484: & \approx & \frac12 \int_0^{t_{{\rm max}}}
485: 	\frac{2}{\pi \sqrt{(n-k)^2-2\ell^2-2(m-k)^2}} \ dk \\
486: & = & \frac{1}{\pi} \left. \left( \tan^{-1} 
487: 	\frac{k+n-2m}{\sqrt{(n-k)^2-2\ell^2-2(m-k)^2}} \right)
488: 	\right|_{k=0}^{k=t_{{\rm max}}} \\
489: & = & \frac12 - \frac{1}{\pi} \tan^{-1}
490: \frac{n-2m}{\sqrt{n^2-2\ell^2-m^2}} \\ 
491: & = & \frac12 + \frac{1}{\pi} \tan^{-1}
492: \frac{2y-1}{\sqrt{1-2x^2-2y^2}}, 
493: \end{eqnarray*}
494: where $k=t_{{\rm max}}$ is the larger of the two roots of the equation
495: $(n-k)^2-2\ell^2-2(m-k)^2=0$, $x=\ell/n$, and $y=m/n$; we truncate the
496: sum and integral at $t_{{\rm max}}$ on the supposition (to be
497: discussed in the next paragraph) that essentially no creation occurs
498: outside the arctic circle.  To make this argument rigorous, we need to
499: deal honestly with the oscillating factor in the creation rate
500: estimate inside the arctic circle, and we need to circumvent the
501: non-uniformity of our estimates.
502: 
503: The non-uniformity of the estimates can be dealt with simply by
504: summing over a smaller interval than in (\ref{crsum}).  The
505: exponentially small bounds on the creation rates outside the arctic
506: circle show that the terms in (\ref{crsum}) that come from locations
507: outside the arctic circle contribute very little to the sum.
508: Motivated by this, we look at the sum of all the terms that come from
509: locations that are far enough inside the arctic circle that our
510: creation rate estimates from \propref{crest} apply.  Because the
511: creation rates are all non-negative, this new sum underestimates the
512: placement probability.  Dealing appropriately with the oscillating
513: factor (as described below) gives an estimate for the new sum; as we
514: see in the computation above, this estimate turns out to be the
515: arctangent formula from \thmref{main}.  A short argument shows that
516: the placement probabilities can be no larger asymptotically
517: (\propref{maininside}), and because our estimate is an underestimate
518: we know they can be no smaller.  This completes the proof.  (Actually,
519: this method works only away from the edges of the diamond, so it is
520: not until Section~\ref{sec-conclusion}~that the proof is completed.)
521: 
522: All that remains is to describe how to deal with the oscillating
523: factor in the summand.  We must show that replacing the oscillating
524: factor by its average value has an asymptotically negligible effect on
525: the sum.  Equivalently, we must show that the difference between the
526: original sum and the smoothed sum is small.  This difference is an
527: exponential sum, and we can estimate it using the Kusmin-Landau
528: Theorem once some preparatory results (Lemmas~\ref{phase},
529: \ref{algebraic}, and \ref{algroots}) are in place.
530: 
531: 
532: \section{Creation Rate Estimates}
533: \label{sec-creation}
534: 
535: Our proof of the asymptotic formula for placement probabilities begins
536: with an estimate of creation rates, which is proved using the saddle
537: point method.  Because \propref{gessel} is most conveniently stated
538: for an Aztec diamond of order $n+1$, we will estimate the creation
539: rates in an Aztec diamond of order $n+1$.  From this point on, we
540: assume that $\ell+m \equiv n \pmod{2}$, because otherwise
541: $\Cr(\ell,m;n+1)$ is necessarily $0$.  As pointed out in
542: subsection~\ref{ssec-outline}, the creation rates behave differently
543: inside and outside the inscribed circle.  If we estimate the creation
544: rates inside it, we get the following result:
545: 
546: \begin{prop}
547: \label{crest}
548: Fix $\varepsilon > 0$.  If $\ell^2 + m^2 \le (1-\varepsilon)n^2/2$ and
549: $\ell+m \equiv n \pmod{2}$, then
550: $$
551: \Cr(\ell,m;n+1) = \frac{4\cos^2
552: \Phi(\ell,m;n)}{\pi\sqrt{n^2-2\ell^2-2m^2}} + O_\varepsilon(n^{-2})
553: $$
554: for some function $\Phi(\ell,m;n)$, which we determine explicitly
555: below. 
556: \end{prop}
557: 
558: The subscript in $O_\varepsilon(n^{-2})$ indicates that the implicit
559: constant depends on $\varepsilon$.  In this paper, if any subscripts
560: appear on a big $O$ term, then the implicit constant depends only on
561: the indicated variables, but the absence of subscripts should not be
562: taken to imply that the implicit constant is absolute.
563: 
564: \begin{proof}
565: Let 
566: \begin{equation}
567: \label{defoff}
568: f(z) = \frac{(1+z)^{n-b}(1-z)^b}{z^a},
569: \end{equation}
570: where $a$ and $b$ are defined as in the statement of \propref{gessel}.
571: To approximate the creation rate, we need to approximate $c(a,b;n)$,
572: which is the constant term of $f(z)$.  The constant term is given by
573: the usual contour integral, which we will approximate using the saddle
574: point method.
575: 
576: Write $a=(1+u)n/2$ and $b=(1+v)n/2$, so that $-1 \le u,v \le 1$.  Note
577: that the $u,v$ coordinates are related to the coordinates in the
578: statement of the proposition by $u=(\ell+m)/n=x+y$ and
579: $v=(\ell-m)/n=x-y$.  We will keep $u$ and $v$ fixed as we send $n$ to
580: infinity.
581: 
582: The critical points of $f(z)$ are
583: $$z_1 = \frac{-v+\sqrt{u^2+v^2-1}}{1-u}$$
584: and
585: $$z_2 = \frac{-v-\sqrt{u^2+v^2-1}}{1-u}.$$
586: Because $\ell^2+m^2<n^2/2$, we have $u^2+v^2<1$.  It follows that
587: $z_1$ and $z_2$ are complex conjugates on the circle $|z|^2=
588: (1+u)/(1-u)$.  We now apply the saddle point method.  To find the
589: constant term of $f(z)$, we integrate $f(z)/(2\pi i \, z)$ about the
590: circle of radius $\sqrt{(1+u)/(1-u)}$ centered at the origin.  One can
591: check that, on this circle, $|f(z)|$ is greatest at the critical
592: points $z_1$ and $z_2$.  (To check it, parametrize the circle by the
593: angle $\theta$ formed with the real axis.  One has $\partial \log
594: |f(z)|^2/\partial \theta = 0$ iff $z$ is one of the two critical
595: points or $z$ lies on the real axis.  At the critical points,
596: $\partial^2 \log |f(z)|^2 / \partial \theta^2 = n(u^2+v^2-1)/(1-v^2) <
597: 0$, so $|f(z)|$ has maxima at these points.  It must have minima on
598: the real axis, since between any two local maxima there must be a
599: local minimum.)  As $n$ goes to infinity, the integral is given
600: asymptotically by the integrals over the parts of the path near the
601: critical points, which can be estimated straightforwardly.  This is
602: the saddle point method.  We will omit the details of the argument
603: leading to the approximation, because they are standard, and can be
604: found, for example, in \cite[pp.~87--89]{asymp}.
605: 
606: The saddle point method tells us that the constant term of $f(z)$ is
607: the sum $Z_1(1+O(n^{-1}))+Z_2(1+O(n^{-1}))$, where
608: \begin{equation}
609: \label{defofz1}
610: Z_1 = \frac{f(z_1)}{2\pi  z_1}
611: \sqrt{\frac{2\pi}{(\log f)''(z_1)}}
612: \end{equation}
613: and 
614: \begin{equation}
615: \label{defofz2}
616: Z_2 = {\overline{Z_1}} = \frac{f(z_2)}{2\pi 
617: z_2}\sqrt{\frac{2\pi}{(\log f)''(z_2)}}.
618: \end{equation} 
619: (For the proof of \propref{crest} we will not need to determine the
620: signs of the square roots in \eqref{defofz1} and \eqref{defofz2}, but
621: they must be chosen so that $Z_1$ and $Z_2$ are complex conjugates.)
622: 
623: Simplifying $z^2(\log f)''(z)$ yields
624: $$
625: z^2(\log f)''(z) = {\frac {n\left
626: (1+u-4\,z^{2}-2\,uz^{2}-4\,vz^{3}+uz^{4}-z^{4}\right )}
627: {2\,(z^2-1)^{2}}}.
628: $$
629: {}From this, one can check that at either critical point of $f(z)$,
630: $z^2(\log f)''(z)$ has absolute value
631: \begin{equation}
632: \label{absofl}
633: |z_i^2(\log f)''(z_i)|
634: =
635: \frac{n}{2}\sqrt{\frac{(1-u^2-v^2)(1-u^2)}{(1-v^2)}}.
636: \end{equation}
637: 
638: Let $\Psi(u,v;n)$ be the phase of $Z_1$, so that $Z_1 = |Z_1|
639: \exp(i\,\Psi(u,v;n)).$ Then
640: $$Z_1+Z_2 = 2|Z_1|\cos\Psi(u,v;n),$$
641: and $c(a,b;n)$ is approximated by
642: \begin{equation}
643: \label{approx}
644: c(a,b;n)= 
645: 2|Z_1|\cos\Psi(u,v;n)+O\left(\frac{|Z_1|}{n}\right) .
646: \end{equation}
647: Of course,
648: \begin{equation}
649: \label{absofz1}
650: |Z_1| = \frac{|f(z_1)|}{2\pi}\sqrt{\frac{2\pi}{|z_1^2(\log
651: f)''(z_1)|}}.
652: \end{equation}
653: Since $|1+z_1|^2 = 2(1-v)/(1-u)$, $|1-z_1|^2=2(1+v)/(1-u)$, and
654: $|z_1|^2=(1+u)/(1-u)$, we see that
655: \begin{equation}
656: \label{absoff}
657: |f(z_1)| =
658: 2^{n/2}\frac{(1-v)^{(n-b)/2}(1+v)^{b/2}}{(1-u)^{(n-a)/2}(1+u)^{a/2}}.
659: \end{equation}
660: 
661: We have $\Cr(\ell,m;n+1) = c(a,b;n)c(b,a;n)/2^n,$ by \propref{gessel}.
662: Interchanging $a$ and $b$ corresponds to interchanging $u$ and $v$.
663: Let $\tilde x$ denote the result of interchanging $u$ and $v$ (and
664: also $a$ and $b$) in the expression $x$, so that, for example,
665: $\widetilde{u-2v} = v-2u$.  When we substitute (\ref{absoff}) and
666: (\ref{absofl}) into (\ref{absofz1}), we see that
667: $$
668: |Z_1||\widetilde{Z_1}| = \frac{2^n}{\pi n \sqrt{1-u^2-v^2}}.
669: $$
670: Hence, by (\ref{approx})
671: $$
672: \Cr(\ell,m;n+1)=
673: \frac{4\cos\Psi(u,v;n)\cos\Psi(v,u;n)}{\pi n
674: \sqrt{1-u^2-v^2}}+O(n^{-2}).
675: $$
676: (To see that the error term is $O(n^{-2})$, one uses the fact that it
677: is $O(|Z_1||\widetilde{Z_1}|/(n2^n))$ and that $|Z_1||\widetilde{Z_1}|
678: = O(2^n/n)$.)
679: 
680: Now we check that $\cos\Psi(v,u;n)=\pm\cos\Psi(u,v;n).$ The identity
681: $$
682: c(b,a;n)b!(n-b)!=c(a,b;n)a!(n-a)!
683: $$
684: suggests that this should be so, but does not seem to prove it.  If we
685: set $\alpha = z_1^2(\log f)''(z_1)$, we find (after some
686: simplification) that
687: $$
688: \frac{\alpha}{\tilde \alpha} = \frac{1-u^2}{1-v^2}.
689: $$
690: Thus, $\alpha$ and $\tilde \alpha$ have the same phase.  If we combine
691: the formulas
692: $$
693: \frac{1+z_1}{1+{\tilde z_1}} = \frac{1-v}{1-u}
694: $$
695: and
696: $$
697: \frac{1-z_1}{1+z_1} = \frac{1+v}{1-v}\,\frac{1}{\tilde z_1}
698: $$
699: with
700: $$
701: f(z_1) =
702: (1+z_1)^n\left(\frac{1-z_1}{1+z_1}\right)^b
703: \left(\frac{1}{z_1}\right)^a,
704: $$
705: we find that $f(z_1)$ equals $\widetilde{f(z_1)}$ times a positive
706: factor, so their phases are equal.  Because
707: $$
708: Z_1 = \pm \frac{f(z_1)}{2\pi}\sqrt{\frac{2\pi}{\alpha}}
709: \quad \hbox{and} \quad
710: \widetilde{Z_1} = \pm
711: \frac{\widetilde{f(z_1)}}{2\pi}\sqrt{\frac{2\pi}{\tilde \alpha}},
712: $$
713: we see that $Z_1$ and $\widetilde{Z_1}$ have the same phase, to within
714: a sign, so $\cos\Psi(v,u;n)=\pm\cos\Psi(u,v;n).$
715: 
716: Finally, we change to the coordinates of our generating function by
717: the substitutions $u=(\ell+m)/n$ and $v=(\ell-m)/n$.  We set
718: $\Phi(\ell,m;n) = \Psi(u,v;n)$.  Then when $\ell+m\equiv n \pmod{2}$,
719: the creation rate at the $(\ell,m)$ location in an Aztec diamond of
720: order $n+1$ is
721: $$\Cr(\ell,m;n+1) =
722: \frac{\pm4\cos^2\Phi(\ell,m;n)}{\pi \sqrt{n^2-2\ell^2-2m^2}}+O(n^{-2}).
723: $$
724: Because creation rates must be non-negative, the $\pm$ sign in this
725: formula can always be taken to be $+$.
726: 
727: The constant implicit in the big $O$ depends continuously on $u$ and
728: $v$.  Thus, for fixed $\varepsilon > 0$ the constant can be chosen
729: uniformly for all $u$ and $v$ with $u^2+v^2\le1-\varepsilon$.  We have
730: therefore proved the result claimed in the statement of the
731: proposition.
732: \end{proof}
733: 
734: Given $\ell$, $m$, and $n$, define
735: \begin{equation}
736: \label{sdef}
737: S_\varepsilon = \{k \in \Z : k \ge 0 \hbox{ and }
738: \ell^2+(m-k)^2\le(1-\varepsilon)(n-k)^2/2 \}.
739: \end{equation}
740: Also, define $k_{{\rm max}}$ to be the greatest element of
741: $S_\varepsilon$, and $k_{{\rm min}}$ to be the least.  (In
742: Section~\ref{sec-placement}, we will sum the creation rates in
743: (\ref{crsum}) as $k$ varies over $S_\varepsilon$.  We will do so to
744: make it possible to apply \propref{crest} to the terms in the sum, as
745: described in subsection~\ref{ssec-outline}.)
746: 
747: \begin{lem}
748: \label{maxbound}
749: Suppose that $|\ell|+|m|
750: \le(1-\delta)n$ for some fixed $\delta > 0$.
751: Then
752: $k_{{{\rm max}}}
753: \le
754: (1-(2-\sqrt{2})\delta + O(\varepsilon))n$.  Hence, 
755: if $\varepsilon$ is small enough compared to $\delta$, then
756: $(n-k_{{{\rm max}}})^{-1} =
757: O(n^{-1}),$
758: and for $k \in S_\varepsilon$ we have
759: $$
760: \frac{1}{\sqrt{(n-k)^2-2\ell^2-2(m-k)^2}} =
761: O(\varepsilon^{-1/2}n^{-1}) 
762:  = O_{\varepsilon,\delta}(n^{-1}).
763: $$
764: \end{lem}
765: 
766: \begin{proof}
767: We have $k_{{{\rm max}}} =
768: (2y-1+\sqrt{2((1-y)^2-x^2)}+O(\varepsilon))n,$ where $x=\ell/n$ and
769: $y=m/n$.  For fixed $y$, this function is clearly maximized at $x=0$.
770: When $x=0$, it becomes a linear function maximized at $y=1-\delta$
771: (for $|x|+|y|\le 1-\delta$).  This yields $k_{{{\rm max}}} \le
772: (1-(2-\sqrt{2})\delta+O(\varepsilon))n$.  Therefore, $(n-k_{{{\rm
773: max}}})^{-1}=O(n^{-1})$ if $\varepsilon$ is small enough compared to
774: $\delta$.
775: 
776: For $k \in S_\varepsilon$, we have
777: $$
778: \ell^2+(m-k)^2 \le (n-k)^2(1-\varepsilon)/2
779: .$$
780: It follows that
781: $$
782: (n-k)^2-2\ell^2-2(m-k)^2 \ge (n-k)^2\varepsilon.
783: $$
784: Therefore,
785: $$
786: \frac{1}{\sqrt{(n-k)^2-2\ell^2-2(m-k)^2}} =
787: O(\varepsilon^{-1/2}(n-k)^{-1})  = O(\varepsilon^{-1/2}n^{-1}).
788: $$
789: If we are not worrying about dependence on $\varepsilon$, this is
790: $O(n^{-1})$.
791: This completes the proof.
792: \end{proof}
793: 
794: In order to apply the Kusmin-Landau Theorem to deal with the
795: exponential sums that will appear later in the proof (as discussed in
796: subsection~\ref{ssec-outline}), we will need to specify $\Phi$, since
797: the phase of $Z_1$ is not uniquely determined.  Also, it will be
798: convenient to extend it to a function of real, and even complex,
799: variables (rather than just integers).
800: 
801: Given a point $(x,y) \ne (0,0)$ in the plane, define $\theta(x,y)$ to
802: be the angle in $(-\pi,\pi]$ formed by the right half of the
803: horizontal axis and the ray from the origin through $(x,y)$.
804: 
805: \begin{lem}
806: \label{phase}
807: We can choose $\Phi(\ell,m;n)$ in \propref{crest} so that 
808: if one sets
809: $\ell=xn$, $m=yn$, and $k={\kappa}n$ in
810: $d\Phi(\ell,m-k;n-k)/dk$, then $d\Phi(\ell,m-k;n-k)/dk$ equals
811: \begin{eqnarray*}
812: \theta\left({-x+y-{\kappa}},
813: {\sqrt {(1-{\kappa})^2-2\,(x^2+(y-{\kappa})^2)}
814: }\right) - \\
815: \theta\left({1-{\kappa}-2\,x},
816: {\sqrt {(1-{\kappa})^2-2\,(x^2+(y-{\kappa})^2)}
817: }\right) + \\
818: {\frac {{x}^{2}-{\kappa}-3\,y{\kappa}+2\,{{\kappa}}^{2}+y+{y}^{2}}
819: {n\sqrt{(1-{\kappa})^2-2\, (x^2+(y-{\kappa})^2)}\,
820: (y+1-2\,{\kappa}-x)(y+1-2\,{\kappa}+x)}}.
821: \end{eqnarray*}
822: As $n$ tends to infinity, the last term is $O(1/n)$ for $k \in
823: S_\varepsilon$ with $\varepsilon > 0$ fixed, as long as $|\ell|+|m|
824: \le (1-\delta)n$ for some fixed $\delta > 0$, and $\varepsilon$ is
825: small enough compared to $\delta$.
826: \end{lem}
827: 
828: \begin{proof}
829: {}From (\ref{defofz1}), we see that we can
830: choose $\Phi(\ell,m;n)$ to be the imaginary part 
831: \begin{equation}
832: \label{choice}
833: \hbox{Im}
834: \left (
835: \log f(z_1)  - \log z_1 - \frac{1}{2}\log((\log f)''(z_1))
836: \right)
837: .
838: \end{equation}
839: 
840: If we substitute $m-k$ for $m$ and $n-k$ for $n$ and differentiate,
841: then the first term of (\ref{choice}) contributes the $\theta$-terms
842: in the formula we are proving.  To see this, recall that (up to an
843: irrelevant multiple of $2\pi i$)
844: $$
845: \log f(z_1) = (n-b)\log(1+z_1) + b \log(1-z_1) - a\log z_1.
846: $$
847: After we express this in terms of $n$, $\ell$, and $m$ and substitute
848: $m-k$ for $m$ and $n-k$ for $n$, the right hand side becomes
849: \begin{equation}
850: \label{messyl}
851: (n/2-k-\ell/2+m/2)\log(1+{\hat z_1}) + 
852: (n/2+\ell/2-m/2) \log(1-{\hat z_1}) \cr
853: - (n/2-k+\ell/2+m/2)\log {\hat z_1 },
854: \end{equation}
855: where $\hat z_1$ is the function of $k$ that results from making the
856: substitutions in $z_1$.  Denote by $L$ the function (\ref{messyl}).
857: When we differentiate $L$ with respect to $k$ (holding $n$, $\ell$,
858: and $m$ fixed), we get
859: $$
860: \frac{\partial L}{\partial k} = 
861: \log{\hat z_1} - \log(1+{\hat z_1}) + 
862: \frac{\partial L}{\partial {\hat z_1}}
863: \frac{\partial {\hat z_1}}{\partial k}.
864: $$
865: Because $z_1$ is a critical point of $f$, ${\partial L}/{\partial
866: {\hat z_1}} = 0$, so ${\partial L}/{\partial k} = \log{\hat z_1} -
867: \log(1+{\hat z_1})$.  Now expressing the imaginary parts of the
868: logarithms in terms of $\theta$ gives the desired terms from the
869: formula we are proving.  (To simplify the terms to the form found in
870: the statement of the lemma, one has to use the fact that for $\alpha >
871: 0$, $\theta(\alpha x,\alpha y) = \theta(x,y)$.)
872: 
873: When we substitute and differentiate, the remaining terms in
874: (\ref{choice}) clearly give algebraic results.  We omit the details of
875: the calculations, since they are tedious and straightforward.
876: 
877: The claim that the last term is $O(1/n)$ for $k \in S_\varepsilon$ is
878: a consequence of \lemref{maxbound}.  The only thing to check is that
879: although the denominator vanishes at ${\kappa} = (y+1\pm x)/2$, these
880: two points are never in $S_\varepsilon$ (or near enough to cause
881: problems).  To see that, note that the definition (\ref{sdef}) of
882: $S_\varepsilon$ is equivalent to the set of $k \ge 0$ for which
883: \begin{equation}
884: \label{othersdef}
885: (1-\kappa)^2-2(x^2+(y-\kappa)^2) \ge
886: \varepsilon(1-\kappa)^2.
887: \end{equation}
888: Note that $\kappa = 1$ is impossible (since then we must have $x=0$
889: and $y=1$, so $|x|+|y| > 1-\delta$).  However, substituting $\kappa =
890: (y+1\pm x)/2$ in the left hand side of (\ref{othersdef}) gives $-(3x
891: \pm(1-y))^2/4$.  Thus, the factors $y+1-2\kappa -x$ and
892: $y+1-2\kappa+x$ in the denominator of the last term in our main
893: formula cannot become arbitrarily small, and the last term is indeed
894: $O(1/n)$.
895: \end{proof}
896: 
897: In Section~\ref{sec-exponential}, we will need the following result
898: (to make it possible to apply exponential sum techniques in the way
899: described in subsection~\ref{ssec-outline}).
900: 
901: \begin{lem}
902: \label{algebraic}
903: The function $d^2 \Phi(\ell,m-k;n-k)/dk^2$ is algebraic.  For any
904: fixed $n$, $\ell$, and $m$ satisfying $|\ell|+|m| < n$ and
905: $\varepsilon > 0$, there exists a neighborhood $U$ in $\C$ of the
906: smallest real interval containing $S_\varepsilon$ such that as a
907: function of $k$, $d^2 \Phi(\ell,m-k;n-k)/dk^2$ is holomorphic on $U$.
908: \end{lem}
909: 
910: \begin{proof}
911: We will use the formula for $d \Phi(\ell,m-k;n-k)/dk$ from
912: \lemref{phase}.  Let $U$ be a small, simply-connected neighborhood in
913: $\C$ of the smallest real interval containing $S_\varepsilon$, such
914: that the points $k=n(y+1\pm x)/2$ are not in $U$.  (We checked at the
915: end of the proof of \lemref{phase} that these points are not in
916: $S_\varepsilon$.)  It follows from the definition of $S_\varepsilon$
917: that
918: $$
919: (n-k)^2-2(\ell^2+(m-k)^2) \ge \varepsilon(n-k)^2 \ge 0
920: $$
921: on $S_\varepsilon$.  If $n=k$, then $\ell = 0$ and $m=k=n$, so
922: $|\ell|+|m| = n$ (contradicting $|\ell|+|m|<n$).  Thus,
923: $(n-k)^2-2(\ell^2+(m-k)^2) \ge \varepsilon$ on $S_\varepsilon$, and
924: hence there is a holomorphic square root of
925: $(n-k)^2-2(\ell^2+(m-k)^2)$ on $U$, if $U$ was chosen to be
926: sufficiently small.  It follows that the third term (the algebraic
927: term) of the formula for $d \Phi(\ell,m-k;n-k)/dk$ in \lemref{phase}
928: is holomorphic on $U$.  The derivative of that term is thus algebraic
929: and holomorphic on $U$, so to complete the proof we just need to check
930: this for the other two terms.
931: 
932: The first two terms can be expressed in terms of the arctangent.  If
933: we do so, we find that the derivative with respect to $k$ of the sum
934: of those two terms is
935: $$
936: \frac{-3x^2+2y\kappa+1-2\kappa-y^2}
937: {n\sqrt{(1-\kappa)^2-2(x^2+(y-\kappa)^2)}
938: (y+1-2\kappa-x)(y+1-2\kappa+x)}.
939: $$ 
940: This is also algebraic and holomorphic on $U$.  Thus,
941: $d^2\Phi(\ell,m-k;n-k)/dk^2$ is holomorphic on $U$ and algebraic, as
942: desired.
943: \end{proof}
944: 
945: We now know everything we need to know about how the creation rates
946: behave inside the arctic circle.  Outside the arctic circle, we can
947: get an exponentially small upper bound for the creation rates.  This
948: will be used for bounding the placement probabilities outside the
949: arctic circle (\propref{arcticbound}).
950: 
951: \begin{prop}
952: \label{expsmall}
953: For each $\varepsilon > 0$, there exists a positive constant $r < 1$
954: such that whenever $\ell^2+m^2 > (1+\varepsilon)n^2/2$,
955: $$
956: \Cr(\ell,m;n+1) = O(r^n).
957: $$
958: \end{prop}
959: 
960: \begin{proof}
961: We assume that $\ell+m\equiv n \pmod{2}$, since otherwise
962: $\Cr(\ell,m;n+1)=0$.  As in the proof of \propref{crest}, we will
963: integrate $f(z)/(2\pi i \, z)$ around a circle about the origin,
964: where, as in (\ref{defoff}),
965: $$
966: f(z) = \frac{(1+z)^{n-b}(1-z)^b}{z^a}.
967: $$
968: However, since we are looking only for an upper bound and not for an
969: asymptotic estimate, we will not need the full saddle point method.
970: We will only sketch the proof, because the details are straightforward
971: but somewhat tedious to check.
972: 
973: We will use the same notation as in the proof of \propref{crest}; for
974: example, we write $a=(1+u)n/2$ and $b=(1+v)n/2$.  Since
975: $u^2+v^2>1+\varepsilon$, the critical points
976: $$
977: z_1 = \frac{-v+\sqrt{u^2+v^2-1}}{1-u}
978: $$ 
979: and
980: $$
981: z_2 = \frac{-v-\sqrt{u^2+v^2-1}}{1-u}
982: $$ 
983: of $f(z)$ are real.  (Of course, the case $u=1$ has to be handled
984: separately, but this will not cause problems.)  We will integrate
985: $f(z)/(2\pi i \, z)$ around a circle of radius $R$, where $R$ will be
986: either $|z_1|$ or $|z_2|$.  We choose $R = |z_i|$ where $|f(z_i)|$ is
987: the lesser of $|f(z_1)|$ and $|f(z_2)|$.  To bound the integral, we
988: will use the fact that the absolute value of the integral is at most
989: as large as the greatest value of $|f(z)|$ on the circle.
990: 
991: It is not hard to check by straightforward manipulation of
992: inequalities that $|f(z_1)| > |f(z_2)|$ if $uv > 0$, and $|f(z_1)| <
993: |f(z_2)|$ if $uv < 0$.  (Since $|u|,|v| \le 1$ and $u^2+v^2 >
994: 1+\varepsilon$, we cannot have $uv=0$.)  Thus, $R=|z_2|$ if $uv >0$,
995: and $R=|z_1|$ otherwise.
996: 
997: Take $i \in \{1,2\}$ so that $R=|z_i|$.  On the circle of radius $R$
998: about $0$, $|f(z)|$ is greatest when $z=z_i$; in fact, the second
999: derivative test shows that this is the only local maximum.  Thus, the
1000: integral is bounded by $|f(z_i)|$, so $|c(a,b;n)| \le |f(z_i)|$.
1001: 
1002: Because the sign of $uv$ doesn't change when $u$ and $v$ are
1003: interchanged, $|\widetilde{f(z_i)}|$ is the lesser of
1004: $|\widetilde{f(z_1)}|$ and $|\widetilde{f(z_2)}|$.  Hence, $|c(b,a;n)|
1005: \le |\widetilde{f(z_i)}|$.  It follows that
1006: $$
1007: \Cr(\ell,m;n+1) \le \frac{|f(z_i)||\widetilde{f(z_i)}|}{2^n}.
1008: $$
1009: 
1010: A simple calculation gives $|f(z_1)||\widetilde{f(z_2)}| =
1011: |f(z_2)||\widetilde{f(z_1)}| = 2^n$.  The inequalities $|f(z_i)| <
1012: |f(z_{3-i})|$ and $|\widetilde{f(z_i)}| < |\widetilde{f(z_{3-i})}|$,
1013: together with the fact that the only dependence on $n$ in any of these
1014: expressions is in the exponent, imply that the creation rate at
1015: $(u,v)$ is $O(r^n)$ for some $r < 1$.  A little more care in the
1016: estimates shows that this bound can be chosen uniformly for $u^2+v^2 >
1017: 1+\varepsilon$, as desired.
1018: \end{proof}
1019: 
1020: \section{Placement Probability Estimates}
1021: \label{sec-placement}
1022: 
1023: Now that we know the creation rates, we can determine the placement
1024: probabilities.  Fix $\delta > 0$.  In this section, we will look at
1025: the placement probabilities $\Pl(\ell,m;n+1)$ at points $(\ell,m)$
1026: satisfying $\ell+m \equiv n \pmod{2}$ and $|\ell| + |m| \le
1027: (1-\delta)n$.  (The congruence condition rules out the placement
1028: probabilities that we know are 0, and the inequality lets us apply
1029: results such as Lemmas~5--7.)
1030: 
1031: {}From (\ref{crsum}), we see that
1032: $$
1033: \Pl(\ell,m;n+1) = \frac{1}{2}\sum_{k \ge 0}{\Cr(\ell,m-k;n+1-k)}.
1034: $$
1035: It will turn out that the creation rates on or beyond the arctic
1036: circle make a vanishing contribution to this sum as $n \rightarrow
1037: \infty$, so we can remove them from the sum without affecting its
1038: asymptotics.  To remove these terms from the sum, fix $\varepsilon >
1039: 0$ (which we assume is small compared to $\delta$, so that we can
1040: apply results such as \lemref{phase}), and look at the sum
1041: $$
1042: \Plt_\varepsilon = \frac{1}{2}\sum_{k \in S_\varepsilon}{\Cr(\ell,m-k;n+1-k)},
1043: $$
1044: where $S_\varepsilon$ is defined by (\ref{sdef}).  (Note that
1045: sometimes $S_\varepsilon = \emptyset$; we will see that this occurs
1046: only when the placement probability is nearly $0$.)  We will
1047: approximate $\Plt_\varepsilon$, and prove that it approximates
1048: $\Pl(\ell,m;n+1)$.  First, we prove a few easy lemmas.
1049: 
1050: \begin{lem}
1051: \label{bound}
1052: Consider the equation $(1-t)^2-2x^2-2(y-t)^2=0$.  For $|x|+|y|<1$,
1053: this equation has two real roots $t$.  The greater root is $0$ iff
1054: $x^2+y^2=1/2$ and $y < 1/2$, and is less than $0$ iff $x^2+y^2>1/2$
1055: and $y < 1/2$.  The lesser root is greater than or equal to $0$ iff
1056: $x^2+y^2 \ge 1/2$ and $y > 1/2$.
1057: \end{lem}
1058: 
1059: \begin{proof}
1060: Since the discriminant of the polynomial is $8(1-x-y)(1+x-y)$, we see
1061: that it has two real roots whenever $|x|+|y|<1$.  Clearly, $0$ is a
1062: root iff $x^2+y^2=1/2$, and since the sum of the roots is $4y-2$, it
1063: is the greater root iff also $y < 1/2$.  One can check the other
1064: claims similarly.
1065: \end{proof}
1066: 
1067: \begin{lem}
1068: \label{estim}
1069: Let $\delta > 0$, and suppose $|x|+|y| \le 1-\delta$.  Let
1070: $\kappa(\varepsilon)$ be any branch of the multivalued function of
1071: $\varepsilon$ defined by $(1-\varepsilon)(1-\kappa)^2 - 2x^2 -
1072: 2(y-\kappa)^2 = 0$.  Then for $\varepsilon > 0$ (and sufficiently
1073: small relative to $\delta$), we have $\kappa(\varepsilon) = \kappa(0)
1074: + O_\delta(\varepsilon)$.  (The constant implicit in the
1075: $O_\delta(\varepsilon)$ does not depend on $x$, $y$, or
1076: $\varepsilon$.)
1077: \end{lem}
1078: 
1079: \begin{proof}
1080: This simply amounts to showing that $\kappa'(\varepsilon)$ is bounded
1081: as a function of $x$, $y$, and $\varepsilon$, for $\varepsilon$
1082: sufficiently small.  If one computes $\kappa(\varepsilon)$ using the
1083: quadratic formula, and then differentiates it with respect to
1084: $\varepsilon$, one finds that it equals a continuous function of $x$,
1085: $y$, and $\varepsilon$ (for $\varepsilon$ near $0$) divided by
1086: $$
1087: \sqrt{(1-\varepsilon)(y-1)^2-(1+\varepsilon)x^2}.
1088: $$
1089: If $\varepsilon$ is small enough compared to $\delta$, then
1090: $\kappa'(\varepsilon)$ will be continuous, and hence bounded, for all
1091: $x$, $y$, and $\varepsilon$ with $|x|+|y|\le 1-\delta$.  Then
1092: $\kappa(\varepsilon)=\kappa(0)+O(\varepsilon)$, as desired.
1093: \end{proof}
1094: 
1095: \begin{prop}
1096: \label{firstapprox}
1097: Let $\delta > 0$ and $\varepsilon > 0$, such that $\varepsilon$ is
1098: sufficiently small compared to $\delta$.  If $|\ell| + |m| \le
1099: (1-\delta)n$ and $\ell+m \equiv n \pmod{2}$, then
1100: $$
1101: {\Plt_\varepsilon} = {\mathcal P}(\ell/n,m/n)
1102: + O_{\varepsilon, \delta}(n^{-1}) + O_\delta(\varepsilon^{1/2}).
1103: $$
1104: \end{prop}
1105: 
1106: \begin{proof}
1107: {}From formula (\ref{crsum}) and \propref{crest}, we see that
1108: $\Plt_\varepsilon$ is approximated by $$\frac{1}{2}\sum_{k \in
1109: S_\varepsilon}{\frac{4\cos^2\Phi(\ell,m-k;n-k)}{\pi
1110: \sqrt{(n-k)^2-2\ell^2-2(m-k)^2}}}+O(n^{-1}).$$ (The error here is
1111: bounded by a constant multiple of $\sum_{j\ge n-k_{{\rm max}}}j^{-2}$,
1112: which is $O((n-k_{{\rm max}})^{-1}),$ and hence $O(n^{-1})$ by
1113: \lemref{maxbound}.)
1114: 
1115: Since $4\cos^2 z = \exp(2iz) + \exp(-2iz) + 2$, we see that
1116: $\Plt_\varepsilon$ is given to within $O(n^{-1})$ by the sum of
1117: \begin{equation}
1118: \label{firstsum}
1119: \frac{1}{2}\sum_{k \in S_\varepsilon}{\frac{2}{\pi
1120: \sqrt{(n-k)^2-2\ell^2-2(m-k)^2}}},
1121: \end{equation}
1122: with
1123: $$
1124: \frac{1}{2}\sum_{k \in S_\varepsilon}{\frac{1}{\pi
1125: \sqrt{(n-k)^2-2\ell^2-2(m-k)^2}}} \exp(2i\,\Phi(\ell,m-k;n-k))
1126: $$
1127: and its complex conjugate.  \propref{expvanish} will show that the
1128: latter two sums are $O(n^{-1})$ as $n$ goes to infinity.
1129: 
1130: Assuming \propref{expvanish}, we can prove the desired limit by
1131: approximating the sum (\ref{firstsum}) with an integral.  The sum is
1132: equal to
1133: \begin{equation}
1134: \label{intap}
1135: \frac{1}{2}\int_{k_{{\rm min}}}^{k_{{\rm max}}} {\frac{2}{\pi
1136: \sqrt{(n-k)^2-2\ell^2-2(m-k)^2}}}\,dk 
1137: +O(n^{-1});
1138: \end{equation}
1139: to see that the error is $O(n^{-1})$, note that the summand (viewed as
1140: a function of a real variable $k$) is $O(n^{-1})$ by
1141: \lemref{maxbound}, and is monotonic on $S_\varepsilon \cap
1142: (-\infty,2m-n)$ and $S_\varepsilon \cap (2m-n,\infty)$.
1143: 
1144: By \lemref{bound}, the polynomial $(n-t)^2-2\ell^2-2(m-t)^2$ has real
1145: roots $t$.  Let $t_{{\rm min}}$ be the lesser root, and $t_{{\rm
1146: max}}$ the greater root.  Then $S_\varepsilon \subset [0,t_{{\rm
1147: max}}]$ if $t_{{\rm max}} \ge 0$, and $S_\varepsilon = \emptyset$ if
1148: $t_{{\rm max}} < 0$.  By \lemref{bound}, we have $t_{{\rm max}}=0$ iff
1149: $\ell^2+m^2=n^2/2$ and $m < n/2$, and $t_{{\rm max}} < 0$ iff
1150: $\ell^2+m^2 > n^2/2$ and $m < n/2$.  In both of these cases, we have
1151: ${\mathcal P}(\ell/n,m/n)=0$ and $\Plt_\varepsilon = O(n^{-1})$.
1152: Thus, we need only deal with the case $t_{{\rm max}} > 0$.
1153: 
1154: Suppose $t_{{\rm max}} > 0$ and $t_{{\rm min}} < 0$, i.e., $\ell^2+m^2
1155: < n^2/2$.  It follows from \lemref{estim} that $k_{{\rm min}} =
1156: O(n\varepsilon)$, and $k_{{\rm max}}=t_{{\rm max}} + O(n\varepsilon)$.
1157: We will approximate the integral in (\ref{intap}) by
1158: \begin{equation}
1159: \label{int2}
1160: \frac{1}{2}\int_{0}^{t_{{\rm max}}} {\frac{2}{\pi
1161: \sqrt{(n-k)^2-2\ell^2-2(m-k)^2}}}\,dk.
1162: \end{equation}
1163: This approximation introduces further error.  To see how large the
1164: error is, first rescale by a factor of $n$, so that the function under
1165: the square root sign becomes $(1-\kappa)^2-2x^2-2(y-\kappa)^2$.
1166: Around a root $r$, this function can be expanded as $-(\kappa-r)^2 \pm
1167: 2\sqrt{2}\sqrt{(y-1)^2-x^2}(\kappa-r)$ (with the sign depending on
1168: which root $r$ is).  Because $|x|+|y|\le 1-\delta$, the coefficient of
1169: $\kappa-r$ cannot become arbitrarily small.  Thus, for small
1170: $\varepsilon$, the error introduced by the approximation is bounded by
1171: a constant (depending on $\delta$) times
1172: $$
1173: \int_0^{\varepsilon} \frac{d\varepsilon'}{\sqrt{\varepsilon'}},
1174: $$
1175: and hence by $O(\varepsilon^{1/2}).$
1176: 
1177: One can evaluate the integral (\ref{int2}) explicitly, because
1178: \begin{equation}
1179: \label{inteval}
1180: \int\!\frac{dk}{\sqrt{(n-k)^2-2\ell^2-2(m-k)^2}}
1181: =
1182: \tan^{-1}\!
1183: \left(\frac{k+n-2m}{\sqrt{(n-k)^2-2\ell^2-2(m-k)^2}}\right)\!.
1184: \end{equation}
1185: As $k \rightarrow t_{{\rm max}}$, the right hand side of
1186: (\ref{inteval}) approaches $\frac\pi 2$ (since the numerator of the
1187: fraction is positive as its denominator approaches $0$).  We see that
1188: (\ref{int2}) evaluates to
1189: $$
1190: \frac{1}{2}+
1191: \frac{1}{\pi}\tan^{-1}
1192: \left(\frac{2m-n}{\sqrt{n^2-2\ell^2-2m^2}}\right).
1193: $$
1194: 
1195: The case with $t_{{\rm min}} \ge 0$ (i.e., $\ell^2+m^2\ge n^2/2$ and
1196: $m > n/2$) is completely analogous, except the integral is over the
1197: interval $[t_{{\rm min}},t_{{\rm max}}]$, rather than $[0,t_{{\rm
1198: max}}]$.  This integral is $1$, so we get that $\Plt_\varepsilon = 1 +
1199: O(n^{-1})+O(\varepsilon^{1/2})$, which agrees with ${\mathcal
1200: P}(\ell/n,m/n)=1$.  This proves the desired result.
1201: \end{proof}
1202: 
1203: We still need to prove that $\Plt_\varepsilon$ approximates
1204: $\Pl(\ell,m;n+1)$.  We do that as follows:
1205: 
1206: \begin{thm}
1207: \label{maininside}
1208: Let $\delta > 0$ and $\varepsilon > 0$, such that $\varepsilon$ is
1209: sufficiently small compared to $\delta$.  If $|\ell|+|m| \le
1210: (1-\delta)n$ and $\ell+m \equiv n \pmod{2}$, then
1211: $$
1212: \Pl(\ell,m;n+1) = {\mathcal P}(\ell/n,m/n) 
1213: + O_\delta(\varepsilon^{1/2})
1214: + O_{\varepsilon,\delta}(n^{-1}).
1215: $$
1216: \end{thm}
1217: 
1218: \begin{proof}
1219: We need to show that $\Plt_\varepsilon$ approximates
1220: $\Pl(\ell,m;n+1)$.  Since \propref{gessel} implies that the creation
1221: rates are all non-negative, and $\Plt_\varepsilon$ is the sum of a
1222: subset of the creation rates appearing in the sum giving
1223: $\Pl(\ell,m;n+1)$, the placement probability must be at least
1224: $\Plt_\varepsilon$.
1225: 
1226: Also, given any point in the Aztec diamond, the north-going placement
1227: probabilities at the four points obtained by rotating it by multiples
1228: of $90^\circ$ about the origin sum to $1$.  This is true because by
1229: rotational symmetry these placement probabilities are equal to the
1230: placement probabilities in each of the four directions at the original
1231: point, which must sum to 1.  This is the content of (\ref{rotation}),
1232: except here it is expressed in terms of the placement probabilities,
1233: rather than the asymptotic formula.
1234: 
1235: One can check by direct computation that ${\mathcal P}(x,y)+{\mathcal
1236: P}(y,-x)+{\mathcal P}(-x,-y)+{\mathcal P}(-y,x)=1.$ If the difference
1237: between $\Plt_\varepsilon$ and the placement probability were not $
1238: O(\varepsilon^{1/2}) + O(n^{-1})$, then the four placement
1239: probabilities would have to sum to more than $1$, which is impossible.
1240: \end{proof}
1241: 
1242: The statement of \thmref{maininside} implies that away from the edges
1243: of the diamond, the placement probabilities converge uniformly.
1244: (Given any $\varepsilon > 0$, the theorem implies that if $n$ is
1245: sufficiently large, then the placement probabilities are within a
1246: constant multiple of $\varepsilon^{1/2}$ of their limiting values.  In
1247: fact, the slightly awkward theorem statement is equivalent to
1248: asserting uniform convergence; we state it that way because it seems
1249: to be the form in which it is most naturally proved, given our setup.)
1250: Thus, assuming \propref{expvanish}, we have very nearly proved
1251: \thmref{main}.  In Section~\ref{sec-conclusion}, we will complete the
1252: proof, using the following proposition:
1253: 
1254: \begin{prop}
1255: \label{arcticbound}
1256: For each $\varepsilon > 0$, there exists a positive constant $r < 1$
1257: such that whenever $\ell^2 + m^2 > (1+\varepsilon)n^2/2$,
1258: $$
1259: \Pl(\ell,m;n+1) =
1260: \begin{cases}
1261: O(r^n)&\hbox{if $m < n/2$, and}\\
1262: 1+O(r^n)&\hbox{if $m > n/2$.}\\
1263: \end{cases}
1264: $$
1265: \end{prop}
1266: 
1267: \begin{proof}
1268: First suppose that $m < n/2$.  The desired result will follow from the
1269: equation
1270: \begin{equation}
1271: \label{probsum}
1272: \Pl(\ell,m;n+1) = \frac{1}{2}\sum_{k \ge 0}{\Cr(\ell,m-k;n+1-k)},
1273: \end{equation}
1274: together with the estimate given by \propref{expsmall}.  First, we
1275: show that \propref{expsmall} applies to the creation rates appearing
1276: in the sum.  Consider
1277: \begin{equation}
1278: \label{simplefn}
1279: \frac{\ell^2+(m-k)^2}{(n-k)^2}
1280: \end{equation}
1281: as a function of $k$.  Its first derivative at $0$ is
1282: $$
1283: 2\frac{\ell^2+m^2-mn}{n^3},
1284: $$
1285: which is greater than $0$ since $\ell^2+m^2 > n^2/2 > mn$.  The only
1286: root of the derivative is
1287: $$
1288: \frac{\ell^2+m^2-mn}{m-n} < 0.
1289: $$
1290: Thus, the function (\ref{simplefn}) is increasing for $0 \le k < n$.
1291: (Note that in (\ref{probsum}) we need only sum up to $k=(m+n)/2$,
1292: since beyond that point $m-k<-(n-k)$ and hence
1293: $\Cr(\ell,m-k;n+1-k)=0$.  Thus, $k$ never reaches the pole in
1294: (\ref{simplefn}) at $n$.)  Therefore, $\ell^2+(m-k)^2 >
1295: (1+\varepsilon)(n-k)^2/2,$ and \propref{expsmall} applies to bound the
1296: creation rates in (\ref{probsum}).
1297: 
1298: Thus, for some constant $s$ between $0$ and $1$,
1299: $$
1300: \Pl(\ell,m;n+1) \le \sum_{k=0}^{(m+n)/2} O(s^{n-k}).
1301: $$
1302: This geometric series is bounded by $O(s^{(n-m)/2}) = O(s^{n/4}).$
1303: This proves the desired bound, with $r=s^{1/4}$.
1304: 
1305: For $m > n/2$, we use the trick of summing the placement probabilities
1306: at the four points obtained by rotating by multiples of $90^\circ$
1307: about the origin.  As in the proof of \thmref{maininside}, the sum
1308: must be $1$, and we know that three of the terms are $O(r^n)$.
1309: Therefore, the fourth must be $1+O(r^n)$, as desired.
1310: \end{proof}
1311: 
1312: \section{Exponential Sums}
1313: \label{sec-exponential}
1314: 
1315: In the proof of \propref{firstapprox}, we needed to show that
1316: $\Plt_\varepsilon$ is within $O(n^{-1})$ of the sum (\ref{firstsum});
1317: to do so, we made use of an estimate whose proof was deferred
1318: (\propref{expvanish}).  In this section, we will derive that estimate.
1319: We begin with the following lemma.
1320: 
1321: \begin{lem}
1322: \label{algroots}
1323: Let $F(x_1,\dots,x_{n+1})$ be an algebraic function of $n+1$ variables
1324: (defined on a subset of $\C^{n+1}$ to be specified shortly), and let
1325: $S$ be a subset of $\C^n$.  Suppose that for each $(y_1,\dots,y_n) \in
1326: S$, there exists an open set $U \subset \C$ such that as a function of
1327: $x_{n+1}$, $F(y_1,\dots,y_n,x_{n+1})$ is (defined and) holomorphic on
1328: $U$.  Then there is a constant $N$ such that for any $(y_1,\dots,y_n)
1329: \in S$, if we regard $F(y_1,\dots,y_n,x_{n+1})$ as a function of
1330: $x_{n+1}$ on the corresponding $U$, then it has at most $N$ roots in
1331: $U$ if it is not identically zero.
1332: \end{lem}
1333: 
1334: \begin{proof}
1335: Since $F(x_1,\dots,x_{n+1})$ is algebraic, it satisfies an equation
1336: \begin{equation}
1337: \label{algeq}
1338: \sum_{i=0}^{d}{p_i(x_1,\dots,x_{n+1})X^i} = 0,
1339: \end{equation}
1340: with $p_0,\dots,p_d$ polynomials (not all identically zero).  Let $N =
1341: \max_i \deg p_i$.  We will show that $N$ has the desired property,
1342: using induction on $n$.
1343: 
1344: We can choose the coefficients $p_i$ so that they have no
1345: (non-constant) common factor.  Fix $y_1 \in \C$, and let $S' =
1346: \{(y_2,\dots,y_n) \in \C^{n-1} : (y_1,\dots,y_n) \in S\}$.  Define
1347: $G(x_2,\dots,x_{n+1}) = F(y_1,x_2,\dots,x_{n+1})$.  Since the
1348: coefficients were taken to have no common factor, they do not all
1349: vanish when we set $x_1=y_1$.  Their degrees do not increase when we
1350: set $x_1=y_1$ (or when we remove common factors), so our lemma follows
1351: by induction on $n$ (applied to $G$ and $S'$), assuming we can prove
1352: it in the case $n=0$.
1353: 
1354: Suppose $n=0$.  Assuming $F$ is not identically zero, we can divide
1355: (\ref{algeq}) by some power of $X$ to get an equation satisfied by $F$
1356: with non-zero constant term, say $p_h(x_1)$.  (A priori, $F$ will
1357: satisfy the new equation only where $F$ is non-zero.  However, since
1358: $F$ is holomorphic on $U$, its zeros are isolated.  By continuity, it
1359: satisfies the equation at its zeros as well as elsewhere.)  Then any
1360: root of $F$ is a root of $p_h$, so $F$ has at most $\deg p_h$ roots,
1361: and hence at most $N$ roots.
1362: \end{proof}
1363: 
1364: \begin{lem}
1365: \label{expbound}
1366: The exponential sums
1367: $$
1368: \sum_{k \in I} \exp(2i\,\Phi(\ell,m-k;n-k))
1369: $$
1370: remain bounded (uniformly in $I$) as $n$ goes to infinity, where $I$
1371: can be any subinterval of $S_\varepsilon$, as long as $|\ell|+|m| \le
1372: (1-\delta)n$ for some fixed $\delta > 0$, and $\varepsilon$ is small
1373: enough compared to $\delta$.
1374: \end{lem}
1375: 
1376: \begin{proof}
1377: To prove this, we will apply the Kusmin-Landau Theorem
1378: (\thmref{kusmin-landau}).  \lemref{algebraic} says that
1379: $d^2\Phi(\ell,m-k;n-k)/dk^2$ satisfies the conditions of
1380: \lemref{algroots}, so there is an absolute upper bound for the number
1381: of roots that it can have as a function of $k$ while $n$, $\ell$, and
1382: $m$ are held fixed (unless it is identically zero for those values of
1383: $n$, $\ell$, and $m$).  Before we apply the Kusmin-Landau Theorem, we
1384: break $S_\varepsilon$ up into a bounded number of subintervals on
1385: which $d\Phi(\ell,m-k;n-k)/dk$ is monotonic.
1386: 
1387: We have to look at the behavior of $d\Phi(\ell,m-k;n-k)/dk$.  As in
1388: \lemref{phase}, set $k={\kappa}n$, $\ell=xn$, and $m=yn$.
1389: \lemref{phase} says that as $n$ goes to infinity,
1390: $d\Phi(\ell,m-k;n-k)/dk$ equals
1391: \begin{equation}
1392: \label{thetaform}
1393: \theta\left({-x+y-{\kappa}},
1394: {\sqrt {(1-{\kappa})^2-2\,(x^2+(y-{\kappa})^2)}
1395: }\right)-\cr
1396: \theta\left({1-{\kappa}-2\,x},
1397: {\sqrt {(1-{\kappa})^2-2\,(x^2+(y-{\kappa})^2)}
1398: }\right)
1399: +O\left(\frac{1}{n}\right).
1400: \end{equation}
1401: We would like to show that when divided by $\pi$, (\ref{thetaform})
1402: stays away from integers.
1403: 
1404: After (\ref{thetaform}) is divided by $\pi$, the only possible
1405: integral values it can take on are $0$, $\pm 1$, and $\pm 2$ (assuming
1406: $n$ is large enough).  If we ignore the $O(1/n)$ term, the rest of the
1407: formula is the difference of the arguments of two points on the same
1408: horizontal line (divided by $\pi$).  Thus, it cannot be $\pm 2$.  It
1409: can be $0$ only if the points coincide or are on the horizontal axis.
1410: It can be $\pm 1$ only if the points are on the horizontal axis.  The
1411: points coincide iff $x+y=1$, which is impossible (since $|\ell|+|m|
1412: \le (1-\delta)n$).  They are on the horizontal axis iff
1413: \begin{equation}
1414: \label{horaxis}
1415: x^2+(y-{\kappa})^2=(1-{\kappa})^2/2.
1416: \end{equation}
1417: The definition (\ref{sdef}) of $S_\varepsilon$ implies that
1418: $$
1419: x^2+(y-{\kappa})^2 \le (1-\varepsilon)(1-{\kappa})^2/2,
1420: $$
1421: so no $k \in S_\varepsilon$ gives a ${\kappa}$ satisfying
1422: (\ref{horaxis}).  (Note that $\kappa = 1$ is impossible since then
1423: $|x|+|y| = |0|+|1| > 1-\delta$.)
1424: 
1425: In fact, the above argument, combined with continuity considerations,
1426: shows that the two points cannot get arbitrarily close to each other
1427: or the horizontal axis, and they clearly cannot get arbitrarily far
1428: {}from the origin.  Thus, even taking into account the $O(1/n)$ term,
1429: $(d\Phi(\ell,m-k;n-k)/dk)/\pi$ really does stay slightly away from
1430: integers as $n \rightarrow \infty$.  Hence, the Kusmin-Landau Theorem
1431: tells us that the exponential sums are bounded (uniformly in $I$).
1432: \end{proof}
1433: 
1434: \begin{prop}
1435: \label{expvanish}
1436: The sum
1437: $$
1438: \sum_{k \in S_\varepsilon} {\frac{1}{\pi
1439: \sqrt{(n-k)^2-2\ell^2-2(m-k)^2}}} \exp(2i\,\Phi(\ell,m-k;n-k)) 
1440: $$
1441: is $O(n^{-1})$ as $n$ goes to infinity, as long as $|\ell|+|m| \le
1442: (1-\delta)n$ for some fixed $\delta > 0$, and $\varepsilon$ is small
1443: enough compared to $\delta$.
1444: \end{prop}
1445: 
1446: \begin{proof}
1447: Let
1448: $$
1449: a(k) = {\frac{1}{\pi \sqrt{(n-k)^2-2\ell^2-2(m-k)^2}}}
1450: $$
1451: and
1452: $$
1453: b(k) = \sum_{k'=k_{{\rm min}}}^{k-1}\exp(2i\,\Phi(\ell,m-k';n-k')).
1454: $$
1455: For $k \in S_\varepsilon$, $a(k) = O(n^{-1})$ (by \lemref{maxbound})
1456: and $b(k)$ is bounded (by \lemref{expbound}).  Suppose $|b(k)| \le B$
1457: for all $k \in S_\varepsilon$.
1458: 
1459: To bound the sum in the statement of the proposition, we will apply
1460: summation by parts.  We have
1461: \begin{eqnarray*}
1462: \sum_{k \in S_\varepsilon} {\frac{\exp(2i\,\Phi(\ell,m-k;n-k))}{\pi
1463: \sqrt{(n-k)^2-2\ell^2-2(m-k)^2}}} 
1464: & = & \sum_{k \in S_\varepsilon}{a(k)(b(k+1)-b(k))} \\
1465: & = & \sum_{k \in S_\varepsilon}a(k)b(k+1) - \sum_{k \in
1466: S_\varepsilon}a(k)b(k) \\ 
1467: & = & \sum_{k=k_{{\rm min}}+1}^{k_{{\rm max}}+1}\!\!a(k-1)b(k) -
1468: \sum_{k=k_{{\rm min}}}^{k_{{\rm max}}}\!a(k)b(k) \\
1469: & = & \sum_{k=k_{{\rm min}}+1}^{k_{{\rm max}}}\!\!b(k)(a(k-1)-a(k)) +
1470: O(n^{-1}).
1471: \end{eqnarray*}
1472: 
1473: This sum is bounded in absolute value by $B \sum_{k}{|a(k-1)-a(k)|} +
1474: O(n^{-1})$.  The function $a(k)$ is monotonic on $(-\infty,2m-n)$ and
1475: $(2m-n,\infty)$ (on the subintervals where it is real, of course), so
1476: within each of these intervals, the sum $\sum_{k}{|a(k-1)-a(k)|}$
1477: telescopes.  The boundary terms are $O(n^{-1})$, and hence the entire
1478: sum is $O(n^{-1})$.
1479: \end{proof}
1480: 
1481: 
1482: \section{Conclusion of the Proof}
1483: \label{sec-conclusion}
1484: 
1485: The results proved in the preceding three sections give us
1486: \thmref{maininside}, a weakened version of Theorem~\ref{main}, in
1487: which we are restricted to estimating the placement probabilities at
1488: normalized locations $(x,y)$ with $|x|+|y|\le 1-\delta$ for some fixed
1489: $\delta>0$.  That is, we are required to keep $(x,y)$ from getting too
1490: close to the boundary of the diamond.  Here we will show how the
1491: restriction on $(x,y)$ can be relaxed, provided that we are careful to
1492: stay away from the points $(\pm\frac12, \frac12)$.
1493: 
1494: Fix $\delta>0$, and consider the region in the Aztec diamond of order
1495: $n$ defined (relative to normalized coordinates) by the constraint
1496: $x^2+y^2>\frac12+\delta$.  This region splits up into four pieces.
1497: \propref{arcticbound} tells us that the north-going placement
1498: probabilities tend uniformly to 1 in the northern piece and to 0 in
1499: the other three pieces.  The only regions that are not covered by this
1500: method are four small curvilinear trapezoids near the points $(\pm
1501: \frac12, \pm \frac12)$, defined by the inequalities
1502: $x^2+y^2\le\frac12+\delta$ and $1-\delta < |x|+|y| \le 1$.  If $(x,y)$
1503: stays away from these four points as $n$ goes to infinity, then we can
1504: indeed conclude that the placement probabilities for north-going
1505: dominos at location $(x,y)$ are as claimed in Theorem~\ref{main}.
1506: This completes the proof of the main theorem, except near the points
1507: $(\pm\frac12,-\frac12)$, which we will now deal with.
1508: 
1509: Let $R$ be the subregion of the Aztec diamond of order $n$ consisting
1510: of the two lower of the four curvilinear trapezoids defined by
1511: $x^2+y^2 \le \frac12 + \delta$ and $1-\delta < |x|+|y| \le 1$.  In
1512: $R$, we use the inequality (\ref{incbound}).  It says that for $h \ge
1513: 0$,
1514: $$
1515: \Pl(\ell,m;n) \le \Pl(\ell,m+h;n+h).
1516: $$
1517: If $n$ is sufficiently large, then for each point $(\ell,m)$ in $R$,
1518: there exists an $h$ such that the point $(\ell,m+h)$ of the diamond of
1519: order $n+h$ has normalized coordinates satisfying $x^2+y^2 \le \frac12
1520: + \delta$, $1-2\delta < |x|+|y| \le 1-\delta$, and $y < 0$.  Let $S$
1521: be the set of all $(x,y)$ satisfying these three constraints.
1522: Inequality (\ref{incbound}) tells us that the placement probabilities
1523: within $R$ are at most as large as those within $S$.  However, the
1524: part of \thmref{main} that we have already proved gives estimates for
1525: the placement probabilities in $S$, and shows that they tend uniformly
1526: to $0$ as $\delta \rightarrow 0$.  (To see the convergence to $0$ most
1527: easily, look at the level curves of the placement probabilities.)  We
1528: thus conclude that as $\delta \rightarrow 0$, the placement
1529: probabilities in $R$ tend uniformly to $0$.  This completes the proof
1530: of \thmref{main}.
1531: 
1532: Unfortunately, our techniques do not give us an explicit bound for the
1533: difference between the placement probabilities and the arctangent
1534: formula in an Aztec diamond of a given order.  This is not because the
1535: methods are inherently ineffective; rather, it is because we have not
1536: determined the dependence on $\varepsilon$ in the $O(n^{-1})$ term of
1537: the error bound in \thmref{maininside}.  To determine it, we would
1538: have to do so for the error term in \propref{crest}, which seems more
1539: trouble than it would be worth (but could perhaps be done).
1540: 
1541: Using these techniques, we can also prove the arctic circle theorem of
1542: \cite{circle}.  One direction, that the regions outside the inscribed
1543: circle are indeed frozen, follows from \propref{arcticbound}.  To see
1544: this, consider the region $R$ defined (relative to normalized
1545: coordinates) by $x^2+y^2>\frac{1}{2}+\varepsilon$, with $\varepsilon >
1546: 0$.  The number of domino spaces in this region of an Aztec diamond of
1547: order $n$ is less than $n^2$, so the probability that a
1548: non-north-going domino will appear in the subregion with $y >
1549: \frac{1}{2}$, or that a north-going domino will appear in the rest of
1550: $R$, is exponentially small, by \propref{arcticbound}.  From this,
1551: we see that with probability approaching $1$ (as $n$ goes to
1552: infinity), all the dominos in $R$ will be aligned in brickwork
1553: patterns, and thus contained in the polar regions.  This is half of
1554: the arctic circle theorem.
1555: 
1556: The other direction, that the polar regions almost never extend
1557: substantially into the interior of the inscribed circle, requires an
1558: additional result for its proof.  Intuitively, it follows from our
1559: main theorem, which tells us that inside the inscribed circle all four
1560: types of placement probabilities are positive.  This trivially implies
1561: that the polar regions cannot almost always cover a given part of the
1562: interior of the circle, but showing that they almost never do is
1563: harder.  We will prove it in subsection~\ref{ssec-ACT}.
1564: 
1565: \section{Consequences of the Theorem}
1566: \label{sec-consequences}
1567: 
1568: \subsection{Height functions}
1569: \label{ssec-height}
1570: 
1571: Height functions for domino tilings, which were introduced in the
1572: mathematics literature in \cite{conway} (and independently in a
1573: slightly different form in the physics literature in \cite{levitov}),
1574: are a very useful device in the study of tilings of simply-connected
1575: subsets of the plane.  (A more general approach to height functions
1576: can be found in \cite{spaces}.)  In any such region that can be tiled
1577: by dominos, the number of enclosed white squares and the number of
1578: enclosed black squares under an alternating coloring of the squares
1579: must clearly be equal.  It follows that if one travels around the
1580: boundary of the region counterclockwise, then one will see a black
1581: square on one's left half the time and a white square on one's left
1582: half the time; to see why, notice that the edges of the square grid
1583: that lie within the region pair sides of black squares with sides of
1584: white squares, so the excess of unpaired (i.e., boundary) sides of
1585: black squares over unpaired sides of white squares is four times the
1586: excess of black squares over white squares.  As one travels around the
1587: boundary, any temporary excess of one kind of square over the other
1588: that is observed along the way represents a ``debt'' that will
1589: eventually have to be paid.  Moreover, the same is true simultaneously
1590: for all the boundaries of all the simply-connected regions that are
1591: formed by suitable subsets of the tiles in question.  Height functions
1592: provide a uniform framework for keeping track of all these debts
1593: simultaneously.
1594: 
1595: If $R$ denotes a finite, simply-connected region composed of lattice
1596: squares that have been alternately colored black and white, a {\sl
1597: height function} on $R$ is an integer-valued function $h$ on the
1598: vertices of the lattice squares which satisfies the following two
1599: properties for adjacent vertices $u$ and $v$: first, if the edge from
1600: $u$ to $v$ is part of the boundary of $R$, then $|h(u)-h(v)|=1$, and
1601: second, if the edge from $u$ to $v$ has a black square on its left,
1602: then $h(v)$ is either $h(u)+1$ or $h(u)-3$.  It is not hard to show
1603: that such a function necessarily satisfies a discrete Lipschitz
1604: condition: if vertices $u$ and $v$ are at distance $d$ from each other
1605: in the sup-norm, then $|h(u)-h(v)| \leq 2d+1$.  Note also that if
1606: $h(\cdot)$ is a height function, then so is $h(\cdot)+C$ for any
1607: integer $C$.
1608: 
1609: Every height function on $R$ determines a domino tiling of $R$,
1610: consisting of dominos that occupy all the domino spaces that are
1611: bisected by edges $uv$ with the property that $|h(u)-h(v)|=3$.
1612: Conversely, every domino tiling of $R$ arises in this way from a
1613: height function on $R$ that is unique modulo addition of a global
1614: constant.  We can remove this ambiguity by constraining a particular
1615: vertex on the boundary of $R$ to have some particular integer value as
1616: its height; then every domino tiling of $R$ has a unique height
1617: function subject to this constraint, and what is more, all these
1618: height functions agree with one another on the boundary of $R$.  For
1619: instance, in the case of the Aztec diamond of order $n$, we set things
1620: up so that the middle vertex on the west edge of the diamond has
1621: height 0 and the middle vertex on the northern edge has height $2n$.
1622: (Note that this differs by 1 from the height function convention used
1623: in \cite{alternating}.)  Then the heights for a typical domino tiling
1624: of an Aztec diamond are as shown in Figure~\ref{fig-typical}.  (The
1625: shading convention for the lattice squares is the same as that in
1626: Section~\ref{sec-introduction}, i.e., so that the leftmost square of
1627: each row in the top half of the diamond is white.)
1628: 
1629: \begin{figure}
1630: \begin{center}
1631: \PSbox{heights.ps}{2.0in}{2.25in}
1632: \end{center}
1633: \caption{A height function for a domino tiling.}
1634: \label{fig-typical}
1635: \end{figure}
1636: 
1637: One can also develop an analogous theory of height functions for other
1638: sorts of tilings, for example, tilings of regions in the triangular
1639: lattice by lozenges (two unit equilateral triangles joined along an
1640: edge).  This theory is simpler geometrically than that for domino
1641: tilings; for the details, see \cite{conway}.  (See also \cite{bh} for
1642: an independent, earlier development of height functions for this
1643: lattice in physics.)  Height functions can furthermore be applied to
1644: the square ice model studied by Lieb, as is shown in \cite{bcsos}.
1645: The results of subsections~\ref{ssec-robustness} and
1646: \ref{ssec-variance} generalize straightforwardly to other sorts of
1647: height functions.  However, because the focus of this article is on
1648: domino tilings, we will not go into the details of the generalization.
1649: 
1650: Suppose $u$, $v$, and $w$ are three consecutive vertices along a path
1651: in a simply-connected region $R$ that is tiled by dominos, such that
1652: neither the edge $uv$ nor the edge $vw$ bisects a domino.  Then
1653: $|h(u)-h(v)|=|h(v)-h(w)|=1$, with $h(w)=h(u)$ if the three points are
1654: collinear and $h(w)=h(u) \pm 2$ if they are not.  This principle makes
1655: it fairly easy to go through the tiling, assigning heights to the
1656: vertices.  Alternatively, one can use this method just to find the
1657: heights along the boundary, and then find the heights in the interior
1658: by the following procedure.  To determine the height of a particular
1659: vertex in the interior of a tiled region, start at the point on the
1660: boundary of the region due north of the vertex (whose height is
1661: independent of the tiling) and proceed downward, subject to the
1662: following rule: when one travels southward along an edge that bisects
1663: a north-going (resp.\ south-going) domino in the tiling, the height
1664: decreases (resp.\ increases) by 3, whereas, when one travels southward
1665: along an edge that bisects a north-going (resp.\ south-going) domino
1666: space that is not occupied by a domino in the tiling, the tiling, the
1667: height increases (resp.\ decreases) by 1.  A similar rule can be
1668: formulated for describing how the height changes as one travels
1669: horizontally through the interior of the diamond.  The fact that these
1670: rules are consistent with each other is a consequence of the fact that
1671: any region that can be tiled by dominos must contain exactly equal
1672: numbers of black and white squares.
1673: 
1674: If we take the average of all the (finitely many) height functions
1675: associated with the different tilings of a region, we get a
1676: real-valued function on the vertices called the {\sl average height
1677: function}.  As a consequence of the rule described in the preceding
1678: paragraph, one can give a simple description of how the average height
1679: changes from vertex to vertex, in terms of the placement probability
1680: $p$ associated with the domino space that is bisected by the edge that
1681: connects the two vertices.  For instance, if $u$ and $v$ are
1682: neighbors, with $u$ to the north or west of $v$, then the average
1683: height at $v$ is equal to the average height at $u$ plus $4p-1$ if
1684: edge $uv$ bisects a south-going or west-going domino space, while the
1685: average height at $v$ is equal to the average height at $u$ minus
1686: $4p-1$ if edge $uv$ bisects a north-going or east-going domino space.
1687: 
1688: Here we are interested in the asymptotic behavior of the average
1689: height function for domino tilings of the Aztec diamond.  The
1690: arctangent formula tells us that these probabilities $p$ are slowly
1691: varying, so the average height function is locally approximated by
1692: functions of the form $ax+by+c$ (with $a$, $b$, and $c$ slowly
1693: varying).  We call the pair $(a,b)$ the {\sl tilt} of the plane
1694: $z=ax+by+c$.  Let us normalize our height functions by dividing
1695: through by $n$, both in the domain and in the range.  Thus, in the
1696: limit we expect to see some sort of function ${\mathcal
1697: H}(\cdot,\cdot)$ on $\{(x,y): |x|+|y| \leq 1\}$ satisfying the
1698: piecewise-linear boundary condition ${\mathcal H}(x,y) = 1-x^2+y^2$
1699: for $|x|+|y|=1$, as well as a Lipschitz condition with constant 2
1700: relative to the sup-norm distance.  In addition, the formulation in
1701: the previous paragraph of how the average height changes when moving
1702: between vertices tells us that we should have $\frac{\partial{\mathcal
1703: H}}{\partial y}=2(p_n-p_s)$ and $\frac{\partial{\mathcal H}}{\partial
1704: x}=2(p_w-p_e)$, where $p_n$, $p_s$, $p_e$, and $p_w$ are the
1705: north-going, south-going, east-going, and west-going placement
1706: probabilities at $(x,y)$, respectively.
1707: 
1708: It can be shown (although we do not prove this here) that the domino
1709: shuffling algorithm of \cite{alternating} can be interpreted directly
1710: in terms of height functions, and that half of the values of the
1711: average height function for the diamond of order $n+1$ are equal to
1712: certain corresponding values of the average height function for the
1713: diamond of order $n$.  Hence the average height functions for the
1714: Aztec diamonds of orders $n$ and $n+1$ cannot be too far apart.
1715: However, such considerations are not sufficient to yield a proof that
1716: the normalized average height functions converge to a continuum limit.
1717: 
1718: The arctangent formula gives us the strength we need in order to
1719: conclude that a limit exists.  Recall that the average height function
1720: can be derived by taking cumulative sums and differences of local
1721: placement probabilities, with various coefficients.  Taking this
1722: assertion to the limit as $n \rightarrow \infty$, and using the fact
1723: that the placement probabilities approach a continuum limit, we see
1724: that the normalized average height function also approaches a limit.
1725: (It is true that the errors in the placement probabilities are going
1726: to add, and that there are more and more of them as $n$ gets large,
1727: but each individual error is going to be small, so that when we divide
1728: by $n$ the normalized error is small as well.)  The limit must be some
1729: function ${\mathcal H}(x,y)$ (defined for $|x|+|y| \leq 1$) with the
1730: property that $\frac{\partial {\mathcal H}}{\partial x}=2{\mathcal
1731: P}(y,-x)-2{\mathcal P}(-y,x)$ and $\frac{\partial {\mathcal
1732: H}}{\partial y}=2{\mathcal P}(x,y)-2{\mathcal P}(-x,-y)$ for all
1733: $(x,y)$ in the interior of its domain.  (Here, we have expressed the
1734: placement probabilities near $(x,y)$ for all four domino orientations
1735: in terms of ${\mathcal P}$ via rotational symmetry.)  That is, the
1736: tilt of the tangent plane at a point (associated with the average
1737: height function) can be expressed in terms of the local placement
1738: probabilities for random domino tilings.  This means that we ought to
1739: be able to reconstruct the function ${\mathcal H}(\cdot,\cdot)$ from
1740: Theorem 1 via integration, making use of the known boundary conditions
1741: satisfied by ${\mathcal H}$.  If we do this, it turns out that
1742: ${\mathcal H}$ can be written in closed form, and indeed, a formula
1743: for ${\mathcal H}$ can be written especially compactly if one makes
1744: use of the formula for ${\mathcal P}(\cdot,\cdot)$ itself.
1745: Specifically, one can verify that the following formula for ${\mathcal
1746: H}(\cdot,\cdot)$ holds:
1747: 
1748: \begin{prop}
1749: \label{heightformula}
1750: The normalized average height functions for large Aztec diamonds
1751: converge uniformly to
1752: $$
1753: {\mathcal H}(x,y) = 2\left(y{\mathcal P}(x,y)-y{\mathcal
1754: P}(-x,-y)+(1-x){\mathcal P}(-y,x)+(1+x){\mathcal P}(y,-x)\right).
1755: $$
1756: \end{prop}
1757: 
1758: \begin{proof}
1759: We simply check that this formula satisfies the differential equations
1760: and boundary conditions.
1761: \end{proof}
1762: 
1763: Within the temperate zone, the average height function is real
1764: analytic; in each of the polar regions, it is an affine function of
1765: $x$ and $y$.  On the arctic circle itself, away from the points $(\pm
1766: \frac12, \pm \frac12)$, the function ${\mathcal H}(x,y)$ is
1767: differentiable but not twice-differentiable.  It takes the value 2 at
1768: the points $(0, \pm 1)$ and the value 0 at the points $(\pm 1,0)$,
1769: with piecewise-linear behavior on the boundary of the normalized
1770: diamond.  The level set ${\mathcal H}=1$ consists of the two line
1771: segments joining midpoints of opposite sides of the normalized
1772: diamond.
1773: 
1774: One may ask, for diamonds of finite order $n$, how closely the
1775: distribution on height functions is clustered around its mean value.
1776: We will see in the next subsection that the standard deviation of the
1777: height at any fixed location in the Aztec diamond of order $n$ is at
1778: most $8\sqrt{n}$.  However, numerical evidence suggests that, at the
1779: center of the diamond, the standard deviation of the height is much
1780: smaller---more like $\log{n}$, or perhaps even less than that.
1781: 
1782: Our formulas for $\frac{\partial {\mathcal H}}{\partial x}$ and
1783: $\frac{\partial {\mathcal H}}{\partial y}$, in combination with the
1784: arctangent formula, yield (within the temperate zone) the equation
1785: \[
1786: \frac{\partial^2 {\mathcal H}}{\partial y^2}
1787: -\frac{\partial^2 {\mathcal H}}{\partial x^2}
1788: = \frac{8}{\pi \sqrt{1-2x^2-2y^2}}.
1789: \]
1790: We can rewrite this equation in a slightly more illuminating way.  For
1791: $t>0$ and $|x|+|y|\le t$, define
1792: $${\overline{\mathcal H}}(x,y,t)=t{\mathcal H}(x/t,y/t).$$
1793: That is, we undo the scaling introduced in Section~\ref{sec-introduction}.
1794: Then off the arctic circle we have
1795: $$
1796: \frac{\partial^2 {\overline{\mathcal H}}}{\partial y^2}-
1797: \frac{\partial^2 {\overline{\mathcal H}}}{\partial x^2}
1798: = 8u(x,y,t),
1799: $$
1800: where
1801: $$
1802: u(x,y,t) = 
1803: \begin{cases}
1804: \frac{1}{\pi\sqrt{t^2-2x^2-2y^2}}&\hbox{if $x^2+y^2<t^2/2$, and}\\
1805: 0&\hbox{if $x^2+y^2\ge t^2/2$.}\\
1806: \end{cases}
1807: $$
1808: This function is a fundamental solution to the wave equation in two
1809: dimensions, with speed of propagation $1/\sqrt{2}$.  That is, $u$ is a
1810: distribution satisfying
1811: $$
1812: \frac{\partial ^2 u}{\partial t^2} = \frac{1}{2}\left(
1813: \frac{\partial^2 u}{\partial x^2} + \frac{\partial^2 u}{\partial y^2}
1814: \right),
1815: $$
1816: $$
1817: u(x,y,0) = 0,
1818: $$
1819: and
1820: $$
1821: \frac{\partial u}{\partial t}(x,y,0) = \delta(x,y),
1822: $$
1823: where $\delta$ is the (two-dimensional) Dirac delta function.  (For
1824: more details on fundamental solutions to the wave equation, see
1825: \cite[p.~164]{rauch}.)
1826: 
1827: Note that \propref{crest} shows that, except for an oscillating
1828: factor, the creation rates also behave like $2u$.  William Jockusch
1829: has shown in personal communication how to use a generating function
1830: developed in \cite{gip} to explain this behavior, by viewing the
1831: creation rates as numerical approximations to a solution of the wave
1832: equation.
1833: 
1834: He has also pointed out that from his methods, one ought to be able to
1835: deduce a weak version of Theorem~\ref{main}.  More specifically, one
1836: should be able to show that in any macroscopic subregion of a randomly
1837: tiled Aztec diamond of order $n$ (i.e., any subregion of size
1838: comparable to that of the diamond), the expected density of
1839: north-going dominos is within $o(1)$ of that predicted by integrating
1840: the arctangent formula; in particular, this would suffice to prove
1841: \propref{heightformula}.  Unfortunately, his methods would not rule
1842: out the possibility of small-scale fluctuations in the placement
1843: probabilities, such as one gets if one looks at placement
1844: probabilities for all horizontal domino spaces rather than just the
1845: north-going ones.
1846: 
1847: \subsection{Robustness}
1848: \label{ssec-robustness}
1849: 
1850: The formula for the average height function that was derived in the
1851: preceding section from the arctangent formula applies not only to
1852: Aztec diamonds, but also to all regions that approximate them in a
1853: suitable sense.  (Here, as hereafter, the term ``region,'' without
1854: qualifiers, should be understood to refer to finite regions in the
1855: plane that are unions of lattice squares and can be tiled with
1856: dominos.)  It is not enough that the region being tiled should have a
1857: boundary that is roughly ``Aztec'' in shape.  For instance,
1858: Figure~\ref{fig-fake1} shows a random domino tiling of a region
1859: obtained from the Aztec diamond of order 32 by adding some squares
1860: along its boundary, while the region that is shown in
1861: Figure~\ref{fig-fake2}, also studied in \cite{remark}, was obtained by
1862: adding an extra row of length 64 in the middle of the diamond.  (These
1863: random tilings were obtained via the method described in \cite{exact}
1864: and are indeed truly random, to the extent that pseudo-random number
1865: generators can be trusted.)  In neither case do we get behavior
1866: consistent with the arctangent formula.  On the other hand,
1867: Figure~\ref{fig-real1} shows a random tiling of an Aztec diamond to
1868: which {\sl two} rows of length 64 have been added, and the resemblance
1869: to Figure~\ref{fig-aztec} is evident.
1870: 
1871: \begin{figure}
1872: \begin{center}
1873: \PSbox{flat.ps}{250pt}{250pt}
1874: \end{center}
1875: \caption{A modified Aztec diamond with far more tilings.}
1876: \label{fig-fake1}
1877: \end{figure}
1878: 
1879: \begin{figure}
1880: \begin{center}
1881: \PSbox{aug32.ps}{250pt}{260pt}
1882: \end{center}
1883: \caption{A modified Aztec diamond with far fewer tilings.}
1884: \label{fig-fake2}
1885: \end{figure}
1886: 
1887: \begin{figure}
1888: \begin{center}
1889: \PSbox{daug32.ps}{250pt}{260pt}
1890: \end{center}
1891: \caption{A modified Aztec diamond with Aztec-like height function.}
1892: \label{fig-real1}
1893: \end{figure}
1894: 
1895: The sense of mystery dissolves if one considers the behavior of the
1896: height function along the boundary in each of the three cases.  In the
1897: first case, the height function is nearly constant along the boundary;
1898: in the second, the direction of change of the height function is the
1899: same along the southwest and northwest edges (and also the same along
1900: the southeast and northeast edges); and in the third, the direction of
1901: change of the height function switches as one rounds any of the four
1902: corners of the region.  Since it is the third situation that resembles
1903: the boundary behavior of height functions of Aztec diamonds, it is not
1904: surprising that the third situation should also give behavior in the
1905: interior that is similar to what one sees for Aztec diamonds.  (In
1906: fact, if we view the third region as an Aztec diamond of order 33 with
1907: two vertical dominos removed, then since almost all tilings of the
1908: Aztec diamond contain those two dominos, the local statistics in the
1909: third region differ little from those in the Aztec diamond of order
1910: $33$.)
1911: 
1912: Note, incidentally, that if instead of adding a row of length 64, as
1913: we did in Figure~\ref{fig-fake2}, we removed a row of length 64, then
1914: the resulting region is easily seen to have a unique tiling,
1915: consisting entirely of horizontal brickwork.  Although this situation
1916: may seem trivial, it can shed some light on what is happening in
1917: Figure~\ref{fig-fake2}.  The horizontal brickwork pattern seen almost
1918: everywhere in Figure~\ref{fig-fake2} is the unique arrangement of
1919: dominos such that the height increases (or decreases, depending on
1920: whether the dominos in the pattern are north-going or south-going) as
1921: quickly as possible as one moves vertically.  In an Aztec diamond with
1922: a row of length 64 removed, the heights on the boundary are such that
1923: the only way to span the gap between the heights on the lower half of
1924: the boundary and those on the upper half is to change at this rate.
1925: Thus, the only way to tile the region is with a brickwork pattern.  In
1926: the case of Figure~\ref{fig-fake2}, the occurrence of an extra row of
1927: length 64 gives the height function a tiny bit of slack, and we can
1928: see where this slack gets used by following the fault-line that runs
1929: {}from left to right.
1930: 
1931: Of course, we could have predicted ahead of time that small
1932: modifications of the shape of the boundary can have a drastic impact
1933: on the tiling situation, since for instance adding a single square to
1934: a region (or removing a single square) can create a region with odd
1935: area, which cannot be tiled at all.  Hence, we will want to assume
1936: that all the regions we discuss actually admit tilings by dominos, as
1937: stipulated in the first paragraph of this subsection.
1938: 
1939: We will show in this subsection that regions similar to Aztec
1940: diamonds, such as Figure~\ref{fig-real1}, have approximately the same
1941: average height functions as the Aztec diamonds they resemble.  This
1942: will follow as a consequence of a more general result, asserting that
1943: the value of the average height function depends in a continuous
1944: manner on the values of the fixed heights along the boundary.  That
1945: is, if one modifies the shape of the boundary in such a way that the
1946: height function along the new boundary, when plotted in three
1947: dimensions (the two original dimensions plus a third dimension for
1948: height), is close to the graph of the old, the average heights of
1949: vertices in the interior will not change very much.
1950: 
1951: Before we can do this, we first prove a general monotonicity result
1952: about height functions.  The idea for this approach was suggested by
1953: Robin Pemantle in personal communication.  Let $R$ denote a
1954: simply-connected region in the plane with some fixed checkerboard
1955: coloring, and let $V$ be the set of vertices in $R$.  Let $V'$ be a
1956: subset of $V$ that contains all the vertices on the boundary of $R$;
1957: we will assume that $V'$ is connected, in the sense that the subgraph
1958: of the square grid induced by the vertex set $V'$ is connected.  A
1959: {\sl partial height function} is a function $f : V' \rightarrow \Z$
1960: subject to the local constraint that if $u$ and $v$ are adjacent
1961: vertices such that the directed edge from $u$ to $v$ has a black
1962: square on its left, then $f(v)-f(u)$ is either $1$ or $-3$.  It is
1963: called {\sl complete height function} if it is defined on all of $V$;
1964: a complete height function $\hat{f}$ {\sl extends} a partial height
1965: function $f$ if it agrees with $f$ where $f$ is defined.
1966: 
1967: Throughout this subsection (and the next), $H$ will denote a complete
1968: height function chosen at random (according to some distribution);
1969: thus, for any vertex $v$, $\Exp[H(v)]$ (the expected value of $H(v)$)
1970: is the value at $v$ of the average height function.
1971: 
1972: Given a connected subset $V'$ of $V$ that contains all the boundary
1973: vertices, and given a partial height function $f$ on $V'$, we let
1974: $\mu_f$ denote the uniform distribution on the set of complete height
1975: functions that extend $f$ to $V$.
1976: 
1977: \begin{lem}
1978: \label{monotone}
1979: If $f$ and $g$ are two partial height functions defined on $V'$ and
1980: agreeing modulo $4$, with $f \leq g$, then $\mu_f$ is stochastically
1981: dominated by $\mu_g$.  That is, there exists a probability measure
1982: $\pi$ on the set of pairs $(\hat{f},\hat{g})$ of complete height
1983: functions extending $f$ and $g$ respectively, such that
1984: $$
1985: \sum_{\hat{g}} \pi(\hat{f},\hat{g})=\mu_f(\hat{f}),
1986: $$
1987: $$
1988: \sum_{\hat{f}} \pi(\hat{f},\hat{g})=\mu_g(\hat{g}),
1989: $$
1990: and
1991: $$
1992: \pi(\{(\hat{f},\hat{g}): \hat{f} \leq \hat{g}\}) = 1.
1993: $$
1994: \end{lem}
1995: 
1996: \begin{proof}
1997: We use induction on the size of $V \setminus V'$ (holding $V$ fixed
1998: and varying $V'$).  The case where this set is empty is trivial.
1999: Assume that the lemma is true whenever $|V \setminus V'| = k-1$, and
2000: suppose we have a situation in which $|V \setminus V'| = k$.  It
2001: clearly suffices to consider the case in which $f(v) < g(v)$ for some
2002: vertex $v$ in $V'$ that is adjacent to at least one vertex in $V
2003: \setminus V'$.  Let $w$ be a vertex in $W = V \setminus V'$ adjacent
2004: to $v$.
2005: 
2006: Given that $f(v)$ has some specific value, any extension of $f$ to
2007: $V'' = V' \cup \{w\}$ would have to give $w$ height $h$ or $h-4$ (for
2008: some particular $h$ whose value we don't care about---it's $f(v)$ plus
2009: or minus 1 or 3), while any extension of $g$ to $V''$ would have to
2010: give $w$ height $h'$ or $h'-4$ (with $h'$ determined from $g(v)$ the
2011: same way $h$ is determined from $f(v)$).  Because $f$ and $g$ agree
2012: modulo $4$ on $V'$ and $h' > h$, we have $h'-4 \ge h$.
2013: 
2014: Let $f'_1$ and $f'_2$ be the two extensions of $f$ to $V''$ that
2015: assign $w$ height $h$ and $h-4$, respectively, and let $g'_1$ and
2016: $g'_2$ be the two extensions of $g$ to $V''$ that assign $w$ height
2017: $h'$ and $h'-4$, respectively.  (If such extensions do not exist, it
2018: is not a problem, as we will see below.)  The distribution $\mu_f$ is
2019: a weighted superposition of $\mu_{f'_1}$ and $\mu_{f'_2}$, where the
2020: $i$th term ($i=1$ or $2$) is given weight proportional to the number
2021: of extensions of $f'_i$ to $V$ (which should be taken to be zero in
2022: the case where the extension to $V''$ does not exist).  Similarly,
2023: $\mu_g$ is a superposition of $\mu_{g'_1}$ and $\mu_{g'_2}$.  Since
2024: $f'_i \leq g'_j$ for all $i,j$ in $\{1,2\}$, and $h \equiv h'
2025: \pmod{4}$, we can use our induction hypothesis to conclude that
2026: $\mu_{f'_i}$ is stochastically dominated by $\mu_{g'_j}$ for all
2027: $i,j$, which implies that $\mu_f$ is stochastically dominated by
2028: $\mu_g$, as was to be shown.
2029: \end{proof}
2030: 
2031: \begin{cor}
2032: \label{concor}
2033: If $f$ and $g$ are two partial height functions on $R$ defined on $V'$
2034: and agreeing modulo $4$, with $f \leq g+4$, then for all $v$,
2035: $\Exp[H(v)]$ under the measure $\mu_f$ is at most 4 more than
2036: $\Exp[H(v)]$ under the measure $\mu_g$.
2037: \end{cor}
2038: 
2039: \begin{proof}
2040: Apply \lemref{monotone} to the partial height functions $f$ and $g+4$.
2041: \end{proof}
2042: 
2043: For applications of \lemref{monotone} and \corref{concor}, it is
2044: important to note that the values of height functions on connected
2045: regions are determined modulo $4$, given the value at any one point,
2046: because the defining conditions for a height function imply that if
2047: two height functions agree modulo $4$ at any point, then they do so at
2048: each neighboring point.  Also, given two partial height functions
2049: defined on different sets, we say that they agree modulo $4$ if all
2050: height functions extending them agree modulo $4$.
2051: 
2052: \begin{prop}
2053: \label{robustness}
2054: Suppose that $R_1$, $R_2$ are two simply-connected regions in the
2055: plane, with mandated partial height functions $f_1$, $f_2$ along their
2056: boundaries that agree modulo $4$, such that every vertex $v_1$ on the
2057: boundary of $R_1$ is within sup-norm distance $\Delta_1$ of some
2058: vertex $v_2$ on the boundary of $R_2$, and vice versa, and such that
2059: whenever vertices $v_1$ and $v_2$ on the respective boundaries are
2060: within sup-norm distance $\Delta_1$ of each other, the heights
2061: $f_1(v_1)$ and $f_2(v_2)$ are within $\Delta_2$ of each other.  Then,
2062: for any $v$ in $R_1 \cap R_2$, the expected value of $H(v)$ under
2063: $\mu_{f_1}$ and the expected value of $H(v)$ under $\mu_{f_2}$ differ
2064: by at most $2\Delta_1 + \Delta_2 + 1$.
2065: \end{prop}
2066: 
2067: \begin{proof}
2068: Let $f_{1,{{\rm max}}}$ be the highest extension of $f_1$ to $R_1$,
2069: let $f_{1,{{\rm min}}}$ be the lowest extension of $f_1$ to $R_1$, and
2070: define $f_{2,{{\rm max}}}$ and $f_{2,{{\rm min}}}$ similarly.  (It is
2071: not hard to show that the complete height functions extending a given
2072: partial height function form a lattice under the usual partial
2073: ordering, so it makes sense to talk about the highest and lowest
2074: extensions.)  Let $v$ be on the boundary of $R_1 \cap R_2$ (and hence
2075: on the boundary of $R_1$ or $R_2$).  If $v$ is on the boundary of
2076: $R_1$, then we can find a nearby $w$ on the boundary of $R_2$ so that
2077: \begin{eqnarray*}
2078: f_{1,{{\rm max}}}(v) & = & f_1 (v) \\
2079: & \leq & f_2 (w) + \Delta_2 \\
2080: & = & f_{2,{{\rm min}}}(w) + \Delta_2 \\
2081: & \leq & f_{2,{{\rm min}}}(v) + 2 \Delta_1 + 1 + \Delta_2,
2082: \end{eqnarray*}
2083: while if $v$ is on the boundary of $R_2$, then we can find a nearby
2084: $w$ on the boundary of $R_1$ so that
2085: \begin{eqnarray*}
2086: f_{1,{{\rm max}}}(v) & \leq & f_{1,{{\rm max}}}(w) + 2 \Delta_1 + 1 \\
2087: & = & f_1(w) + 2 \Delta_1 + 1 \\
2088: & \leq & f_2(v) + \Delta_2 + 2 \Delta_1 + 1 \\
2089: & = & f_{2,{{\rm min}}}(v) + \Delta_2 + 2 \Delta_1 + 1.
2090: \end{eqnarray*}
2091: Since the two height functions agree modulo $4$ at $v$, $f_{1,{{\rm
2092: max}}} (v) \leq f_{2,{{\rm min}}} (v) + 4K$, where $4K$ is the
2093: greatest multiple of 4 that is less than or equal to $2\Delta_1 +
2094: \Delta_2 + 1$.  It follows from this (and the corresponding inequality
2095: in the other direction) that if $f_1'$ is any extension of $f_1$ to
2096: $R_1$ and $f_2'$ any extension of $f_2$ to $R_2$, then for each $v$ on
2097: the boundary of $R_1 \cap R_2$, $f_1'(v)$ differs from $f_2'(v)$ by at
2098: most $4K$.
2099: 
2100: Now let $v$ be any vertex in $R_1 \cap R_2$.  If we compute
2101: $\Exp[H(v)]$ by conditioning on the heights on the boundary of $R_1
2102: \cap R_2$, then it follows from \corref{concor} that the expected
2103: value of $H(v)$ under $\mu_{f_1}$ differs by at most $4K$ (and hence
2104: at most $2\Delta_1 + \Delta_2 + 1$) from its expected value under
2105: $\mu_{f_2}$.
2106: \end{proof}
2107: 
2108: As an application of this result, we may consider a modification of
2109: the Aztec diamond of order $n$, whose symmetric difference with the
2110: true Aztec diamond of order $n$ is a narrow fringe around the border
2111: of the true diamond, of width $o(n)$.  Suppose that the black and
2112: white squares of the symmetric difference are equinumerous, and
2113: moreover that they are not segregated but intermix in such a manner
2114: that the height function along the border of the modified diamond is
2115: within $o(n)$ of the height function along the border of the true
2116: diamond.  Lastly, let us suppose that the modified diamond has at
2117: least one domino tiling.  Then we can conclude that the average height
2118: function for the modified diamond is within $o(n)$ of the average
2119: height function for the true diamond.
2120: 
2121: Notice that these results give us no direct information about how
2122: individual placement probabilities change in response to small changes
2123: in the shape of the boundary, though some weak information can be
2124: obtained by way of the height function.  It would be quite interesting
2125: to obtain robustness results for the placement probabilities
2126: themselves.
2127: 
2128: 
2129: \subsection{Variance}
2130: \label{ssec-variance}
2131: 
2132: In \propref{heightformula} of subsection~\ref{ssec-height}~we gave a
2133: formula for the normalized average height function, or rather its
2134: limit as the size $n$ of the Aztec diamond goes to infinity, and in
2135: subsection~\ref{ssec-robustness}~we showed that the same formula
2136: applies to many regions that are roughly similar to the Aztec diamond.
2137: However, we obtained no information about how closely a typical height
2138: function for a region (an Aztec diamond or something else) should
2139: approximate the average height function.  Here we use Azuma's
2140: Inequality \cite[p.~85]{method} to bound the amount of variation that
2141: values of random height functions are likely to exhibit.
2142: 
2143: Let $H$ denote the (unnormalized) height function corresponding to a
2144: random domino tiling of some simply-connected region in the plane, and
2145: let $v$ denote a vertex in the region, such that there is a path of
2146: $m$ vertices from the boundary of the region to $v$.  We will show in
2147: this subsection that the variance of the random variable $H(v)$ is at
2148: most $64m$.  In fact, we actually get a stronger result:
2149: 
2150: \begin{thm}
2151: \label{variancethm}
2152: Let $f$ be a partial height function defined on the boundary of a
2153: simply-connected region $R$, and let $v$ be a vertex in the interior
2154: of $R$, such that there is a path of $m$ vertices from the boundary of
2155: $R$ to $v$.  Then, for all $c>0$, the probability that $H(v)$ (the
2156: value of a random height function at $v$ under the uniform
2157: distribution $\mu_f$) differs from its expected value by more than $c
2158: \sqrt{m}$ is less than $2e^{-c^2 / 32}$.
2159: \end{thm}
2160: 
2161: \begin{proof}
2162: Let $x_0, x_1, \dots, x_{m-1} = v$ be a lattice-path connecting a
2163: point $x_0$ on the boundary of $R$ to the point $v$.  Let $F_k$ be the
2164: partition of the space of possible height functions in which two
2165: height functions are regarded as equivalent if they agree at
2166: $x_0,x_1,\dots,x_{k-1}$.  Let $M_k$ be the conditional expectation
2167: $\Exp[H(v) | F_k]$, the function from the set of height functions to
2168: the reals that assigns to each height function $h$ the average value
2169: of $h'(v)$ as $h'$ ranges over all height functions in the equivalence
2170: class of $h$.
2171: 
2172: Note that $M_m$ is just the function $H(v)$ itself, while $M_0$ is the
2173: average value of the height at $v$, averaged over all height
2174: functions.  The functions $M_0, M_1, \dots, M_m$ form a martingale;
2175: that is,
2176: \[
2177: \Exp[M_{k+1} | F_k] = M_k.
2178: \]
2179: On each component of $F_k$, $M_{k+1} = \Exp[H(v) | F_{k+1}]$ takes on
2180: at most two distinct values, according to the two different values of
2181: $H(x_k)$ that are consistent with the already-known values of
2182: $H(x_0),H(x_1),\dots,H(x_{k-1})$.  From \corref{concor}, we see that
2183: these two values of $M_{k+1}$ differ by at most 4.  Since $M_k$ is
2184: their weighted average, it follows that $M_{k}$ and $M_{k+1}$ never
2185: differ by more than 4.  Then, applying Azuma's Inequality
2186: (Corollary~2.2 on page~85 of \cite{method}) to the quantities $M_k/4$,
2187: we get
2188: \[
2189: \Prob[|M_m-M_0|/4 > t \sqrt{m} ]  <  2e^{-t^2 / 2}.
2190: \]
2191: Replacing $t$ by $c/4$, we get
2192: \[
2193: \Prob[|H(v)-\Exp[H(v)]| > c \sqrt{m} ]  <  2e^{-c^2 / 32}.
2194: \]
2195: This completes the proof.
2196: \end{proof}
2197: 
2198: If we are interested in estimating the variance, we can derive a
2199: consequence of the preceding inequality: Assuming for simplicity of
2200: derivation (and without loss of generality) that $\Exp[H(v)]=0$, we
2201: have
2202: \begin{eqnarray*}
2203: \Var[H(v)] & = & \Exp[(H(v))^2] \\
2204: & = & \int_0^\infty \Prob[(H(v))^2 > x] \ dx \\
2205: & = & \int_0^\infty \Prob[|H(v)| > \sqrt{x}] \ dx \\
2206: & < & \int_0^\infty 2e^{-x/(32m)} dx \\
2207: & = & 64m.
2208: \end{eqnarray*}
2209: 
2210: As a final aside, we mention that our proof of \thmref{variancethm}
2211: also yields the following more general result:
2212: \begin{prop}
2213: \label{large}
2214: Let $R$ be a simply-connected region in the plane and let $v,w$ be
2215: vertices in the interior of $R$, such that there is a path of $m$
2216: vertices from $v$ to $w$, staying entirely within $R$.  Then, for all
2217: $c>0$, the probability that $H(v)-H(w)$ (under the uniform
2218: distribution on domino tilings of $R$) differs from its expected value
2219: by more than $c \sqrt{m}$ is less than $2e^{-c^2 / 32}$.
2220: \end{prop}
2221: 
2222: \subsection{The arctic circle theorem}
2223: \label{ssec-ACT}
2224: 
2225: We will now use \thmref{variancethm} to complete the proof of the
2226: arctic circle theorem, which we began in Section~\ref{sec-conclusion}.
2227: We still need to show that the polar regions almost never extend very
2228: far into the interior of the inscribed circle $x^2+y^2=\frac12$
2229: (defined relative to normalized coordinates).  Let $\varepsilon > 0$,
2230: and consider the region $R$ in an Aztec diamond of order $n$ defined
2231: by $x^2+y^2<\frac12-\varepsilon$.
2232: 
2233: A domino is in the north or south polar region if and only if the
2234: heights on the vertices of the domino are equal to those at the same
2235: locations in the all-horizontal tiling, which is the minimal tiling of
2236: the Aztec diamond (under the partial ordering of tilings induced by
2237: comparison of height functions).  An analogous statement connects the
2238: other two polar regions to the all-vertical tiling, which is the
2239: maximal tiling.  \propref{heightformula} shows that, asymptotically,
2240: the average height function disagrees with the minimal and maximal
2241: height functions within the inscribed circle (although outside of that
2242: circle it agrees with one or the other).  In particular, if a domino
2243: in $R$ is part of the polar regions, then the heights on it differ
2244: {}from the average heights at those locations by an amount at least
2245: proportional to $n$.  (Of course, the constant of proportionality
2246: depends on $\varepsilon$.)  We see that, by taking $c=\sqrt{n}$ in
2247: \thmref{variancethm}, the probability that a domino in $R$ will be
2248: part of the polar regions is exponentially small in $(\sqrt{n})^2=n$.
2249: Since the number of dominos in $R$ is on the order of $n^2$, the
2250: probability that any will be contained in the polar regions is
2251: exponentially small.
2252: 
2253: We have now proved a slightly stronger version of the arctic circle
2254: theorem than that proved in \cite{circle}.  There, it is shown that
2255: for any $\varepsilon > 0$, for sufficiently large $n$, the boundary of
2256: the temperate zone stays within distance $\varepsilon n$ of the arctic
2257: circle with probability greater than $1-\varepsilon$.  We have shown
2258: that this probability differs from $1$ by an amount exponentially
2259: small in $n$.
2260: 
2261: \subsection{Heterogeneity}
2262: \label{ssec-heterogeneity}
2263: 
2264: The arctangent formula gives us an indication of a certain sort of
2265: local homogeneity: places in the tiling that are close together tend
2266: to be governed by the same statistics.  Here we would like to prove a
2267: converse result, and show that within the temperate zone, places in
2268: the tiling that are far apart tend to be governed by different
2269: statistics.  More precisely, we would like to show that within the
2270: temperate zone, the quadruple
2271: $$
2272: (p_n,p_e,p_s,p_w) = ({\mathcal P}(x,y),{\mathcal P}(-y,x),{\mathcal
2273: P}(-x,-y),{\mathcal P}(y,-x))
2274: $$
2275: (whose components are the four placement probabilities near the
2276: location $(x,y)$) uniquely determines $x$ and $y$.  This means, to put
2277: it somewhat fancifully, that if you found yourself stranded somewhere
2278: in the temperate zone of a random domino tiling of a huge checkerboard
2279: colored Aztec diamond, then, provided that you had a compass to tell
2280: you which way was north, you could determine your relative position
2281: within the diamond merely by examining the local statistics of the
2282: tiling.
2283: 
2284: The heterogeneity claim is not hard to prove, since we know that the
2285: level sets for all four placement probabilities are arcs of ellipses
2286: having a very specific geometry.  In particular, level sets for $p_n$
2287: and $p_s$ are arcs of ellipses that intersect in at most two points,
2288: and these two points have the same $y$-coordinate; similarly, level
2289: sets for $p_e$ and $p_w$ are arcs of ellipses that intersect in at
2290: most two points, and these two points have the same $x$-coordinate.
2291: It follows that if two points have the same probability quadruples,
2292: they must have the same $x$- and $y$-coordinates; that is, the two
2293: points must coincide.
2294: 
2295: However, we wish to prove more.  Consider that the elements of the
2296: quadruple sum to 1, so that the quadruple has three degrees of
2297: freedom.  However, $x$ and $y$ together embody only two degrees of
2298: freedom, so as $x$ and $y$ sweep through their range of joint allowed
2299: values, the quadruple determined by $x$ and $y$ will not sweep through
2300: the full set of probability vectors of length 4.  On the other hand,
2301: the asymptotic normalized average height function ${\mathcal H}$
2302: introduced in subsection~\ref{ssec-height}~manifests exactly two of
2303: the degrees of freedom of $(p_n,p_e,p_s,p_w)$ in its first-order
2304: derivatives.  What we would like to show is that inside the temperate
2305: zone, the map $(x,y) \mapsto (\frac{\partial {\mathcal H}}{\partial
2306: x}, \frac{\partial {\mathcal H}}{\partial y})$ is one-to-one, and has
2307: as its range the region $\{(s,t): |s|+|t| < 2\}$.  (The possible
2308: significance of this fact will be explained more fully in
2309: Section~\ref{sec-speculations}.)  Putting it differently, we may say
2310: that if one views the graph of the restriction of the function
2311: ${\mathcal H}$ to the interior of the temperate zone as a surface,
2312: then the Gauss map from the surface to the sphere is injective.
2313: 
2314: To prove the claim, we first note that, as discussed in
2315: subsection~\ref{ssec-height}, $\frac{\partial {\mathcal H}}{\partial
2316: x}=2p_w-2p_e$ and $\frac{\partial {\mathcal H}}{\partial
2317: y}=2p_n-2p_s$.  With $y$ fixed and $x$ increasing, $p_e$ increases
2318: while $p_w$ decreases, achieving equality (by symmetry) at $x=0$.
2319: Thus, the sign of $\frac{\partial {\mathcal H}}{\partial x}$ tells us
2320: the sign of $x$, and similarly, the sign of $\frac{\partial {\mathcal
2321: H}}{\partial y}$ tells us the sign of $y$.  Hence, to prove the
2322: injectivity of the map, it suffices to focus on the part of the
2323: temperate zone that lies in the interior of one particular quadrant,
2324: say the second.  Within that quarter-disk, $\frac{\partial {\mathcal
2325: H}}{\partial x}$ and $\frac{\partial {\mathcal H}}{\partial y}$ are
2326: both non-negative functions, taking the values $0$ on the respective
2327: axes $x=0$ and $y=0$ and increasing as one moves away from these axes.
2328: These monotonicity properties do not of themselves rule out the
2329: possibility that $\frac{\partial {\mathcal H}}{\partial x}$ and
2330: $\frac{\partial {\mathcal H}}{\partial y}$ have the same value for two
2331: different points in that quarter-disk, so we must resort to a slightly
2332: more arduous approach.
2333: 
2334: Using the arctangent formula, one can check that
2335: $$
2336: \frac{\cos\Big(\frac{\pi}{2}\,\frac{\partial{\mathcal H}}{\partial
2337: y}\Big)}
2338: {\cos\Big(\frac{\pi}{2}\,\frac{\partial{\mathcal H}}{\partial x}\Big)}
2339:  = \frac{1-x^2-3y^2}{1-3x^2-y^2}
2340: % The \Big's are used here instead of \left and \right in order to
2341: % ensure that the parentheses in the numerator and denominator have
2342: % the same size.
2343: $$
2344: and
2345: $$
2346: \frac{\sin\Big(\frac{\pi}{2}\,\frac{\partial{\mathcal H}}{\partial
2347: y}\Big)} 
2348: {\sin\Big(\frac{\pi}{2}\,\frac{\partial{\mathcal H}}{\partial x}\Big)}
2349: = -\frac{y}{x}.
2350: % See the note above.
2351: $$
2352: (If $3x^2+y^2=1$, then the first ratio is not defined.  However, since
2353: $x^2+y^2<\frac12$, either the first ratio or its reciprocal is
2354: defined.)  Given the values of these two ratios, there are in general
2355: at most two possibilities for $(x,y)$, only one of which will be in
2356: the desired quadrant.  The only case in which knowledge of the two
2357: ratios does not restrict us to at most two possibilities for $(x,y)$
2358: is when the first ratio is $1$.  This happens iff $\frac{\partial
2359: {\mathcal H}}{\partial x} = \frac{\partial {\mathcal H}}{\partial y}$,
2360: i.e., along the line through the origin that bisects the quadrant.
2361: Since one can check using the explicit formulas for $\frac{\partial
2362: {\mathcal H}}{\partial x}$ and $\frac{\partial {\mathcal H}}{\partial
2363: y}$ that the partial derivatives increase as one moves away from the
2364: axes along that line, they still determine $(x,y)$.  It follows that
2365: the map $(x,y) \mapsto (\frac{\partial {\mathcal H}}{\partial
2366: x},\frac{\partial {\mathcal H}}{\partial y})$ is injective on the
2367: quadrant, as was to be shown.
2368: 
2369: Now we will see that the map is in fact a surjection to the set
2370: $\{(s,t): |s|+|t|<2\}$.  If one sets $x=(1-t-ct^2)/2$ and
2371: $y=(1+t-ct^2)/2$ with $c > \frac12$ (so that $(x,y)$ lies on a
2372: parabola that is symmetric about the axis $x=y$ and that lies inside
2373: the closed temperate zone in the vicinity of $(\frac12,\frac12)$),
2374: then, sending $t$ to zero from above, we find that the north-going and
2375: east-going probabilities tend towards
2376: \[
2377: \frac{1}{2} + \frac{1}{\pi} \tan^{-1}\frac{1}{\sqrt{2c-1}}
2378: \]
2379: and 
2380: \[
2381: \frac{1}{2} + \frac{1}{\pi} \tan^{-1}\frac{-1}{\sqrt{2c-1}} , 
2382: \]
2383: which sum to 1 for all $c$ between $\frac12$ and infinity and which
2384: vary (as an ordered pair) over the open segment connecting $(1,0)$ to
2385: $(\frac12,\frac12)$, as $c$ goes from $\frac12$ to infinity.  Plugging
2386: the limits $p_n \rightarrow p$, $p_s \rightarrow 0$, $p_e \rightarrow
2387: 1-p$, $p_w \rightarrow 0$ into the formulas $\frac{\partial {\mathcal
2388: H}}{\partial x}=2p_w-2p_e$ and $\frac{\partial {\mathcal H}}{\partial
2389: y}=2p_n-2p_s$, we find that the boundary of the open square $\{(s,t):
2390: |s|+|t|<2\}$ consists of limit points of the set of tilts
2391: $(\frac{\partial {\mathcal H}}{\partial x}, \frac{\partial {\mathcal
2392: H}}{\partial y})$ that are achieved by the average height function in
2393: the temperate zone, and hence (by continuity) that that we do indeed
2394: obtain the open square as the set of tilts achieved by the average
2395: height function in the temperate zone.
2396: 
2397: \subsection{Entropy}
2398: \label{ssec-entropy}
2399: 
2400: The {\sl entropy} of a random variable that takes on any of $N$ values
2401: with respective probabilities $q_1,\dots,q_N$ is defined as
2402: $\sum_{i=1}^N -q_i \log q_i$ (with $0 \log 0 = 0$ by convention); for
2403: example, the entropy of a uniform random domino tiling of the Aztec
2404: diamond of order $n$ is $\frac{n(n+1)}{2} \log 2$, because there are
2405: exactly $2^{n(n+1)/2}$ tilings (see \cite{alternating} for a proof).
2406: We have seen that for large $n$, nearly all of this entropy is due to
2407: the variety exhibited inside, as opposed to outside, the arctic
2408: circle.  It would be good to have more quantitative information on
2409: this.  Specifically, given a patch of an Aztec diamond, one can define
2410: a random variable whose values are the near-tilings of the patch that
2411: result from restricting a uniform random tiling of the Aztec diamond
2412: to just the patch (such near-tilings are allowed to have untiled
2413: squares along the boundary of the patch), and one can consider the
2414: entropy of this new random variable.  If the patch is very large
2415: (while not long and skinny, for example like a $2 \times n$
2416: rectangle), but the order of the Aztec diamond is much larger still,
2417: then we believe that this entropy, when divided by the area of the
2418: patch, is close to a value which we would call the {\sl local
2419: entropy}, and which would depend only on the normalized location of
2420: the patch.
2421: 
2422: In this subsection, we make a small start towards calculating local
2423: entropy by showing that it vanishes outside the arctic circle and that
2424: it is positive inside the arctic circle (assuming it is well-defined
2425: there).  Assuming that local entropy is well-defined everywhere, this
2426: gives us another way of interpreting the arctic circle, namely as the
2427: boundary between the zero-entropy region and the positive-entropy
2428: region.
2429: 
2430: The vanishing (and perforce the well-definedness) of local entropy
2431: outside the temperate zone is a simple consequence of the arctic
2432: circle theorem.  To prove the other half of our claim, consider an $m
2433: \times m$ patch $P$ sitting inside the temperate zone of an extremely
2434: large Aztec diamond, with $m$ even.
2435: 
2436: If $a$, $b$, $c$, and $d$ are the northwest, northeast, southwest, and
2437: southeast squares in a $2\times 2$ block in a plane region $R$ that
2438: can be tiled by dominos, then the proportion of tilings of $R$ that
2439: have a horizontal domino covering squares $a$ and $b$ and another
2440: horizontal domino covering squares $c$ and $d$ (write this proportion
2441: as $p_{ab,cd}$ for short) is clearly equal to the proportion
2442: $p_{ac,bd}$ of tilings that contain vertical dominos covering squares
2443: $a$ and $c$ and squares $b$ and $d$; moreover, by one of the lemmas
2444: proved in \cite{gip}, both proportions are equal to
2445: $p_{ab}p_{cd}+p_{ac}p_{bd}$, where $p_{ab}$ denotes the proportion of
2446: tilings that have a domino covering $a$ and $b$, etc.\ (that is,
2447: $p_{ab}$, $p_{cd}$, $p_{ac}$, and $p_{bd}$ are just placement
2448: probabilities under uniform random tiling).  In our particular
2449: situation, if one looks inside the patch $P$ taken from the temperate
2450: zone of a large Aztec diamond, all four placement probabilities are
2451: bounded away from zero, say by $\varepsilon>0$, so the probability
2452: that a random tiling contains a $2 \times 2$ block centered at any
2453: particular vertex in $P$ is at least $4\varepsilon^2$.  In particular,
2454: we can look at the $(m/2)^2$ vertices that are at the centers of the
2455: $(m/2)^2$ non-overlapping $2 \times 2$ blocks into which $P$ can be
2456: naturally decomposed.  Using linearity of expectation, we can see that
2457: the expected number of such $2\times 2$ blocks in a random tiling of
2458: the Aztec diamond is at least $m^2 \varepsilon^2$.  However, this
2459: allows us to set a lower bound on the entropy, as measured by the
2460: variety of configurations one sees locally.  For, by freely rotating
2461: these blocks (i.e., changing horizontal blocks to vertical blocks or
2462: vice versa), we can create $2^{m^2 \varepsilon^2}$ other local
2463: patterns, all equally likely.  Standard techniques in information
2464: theory permit one to conclude that the entropy of the near-tiling of
2465: $P$ is at least $\varepsilon^2 \log 2$ times the area of $P$.
2466: 
2467: \section{Further Results}
2468: \label{sec-further}
2469: 
2470: Although we have phrased our results in terms of domino tilings, there
2471: is an easy equivalence between domino tilings of finite regions and
2472: dimer configurations on certain finite graphs.  Specifically, if we
2473: replace each square cell by a vertex, and draw an edge connecting any
2474: two vertices whose associated cells are adjacent, then a domino tiling
2475: of a region corresponds to a dimer-cover of the derived graph, that
2476: is, to a set of edges of the derived graph with the property that
2477: every vertex belongs to exactly one of the chosen edges.  In this way,
2478: the study of domino tilings is seen to be equivalent to the study of
2479: dimer-covers, which is one of the better-understood statistical
2480: mechanics models in two dimensions.  Studying domino tilings of
2481: special regions, such as Aztec diamonds, is tantamount to studying the
2482: dimer model in the presence of special boundary conditions.  The
2483: uniformity of the distribution corresponds to a degenerate situation
2484: in which all dimer configurations have the same energy.
2485: 
2486: There has been surprisingly little work on the behavior of the dimer
2487: model in the presence of general boundary conditions; researchers in
2488: statistical mechanics have tended to study either toroidal (i.e.,
2489: periodic) boundary conditions or boundary conditions that correspond
2490: to domino tilings of a rectangle.  Our work can in a sense be regarded
2491: as a somewhat strange chapter in the study of the dimer model, in
2492: which highly unphysical boundary conditions are imposed.  (Precursors
2493: of this research include \cite{elser}, \cite{gg}, and \cite{remark}.)
2494: 
2495: In his original article on the dimer model \cite{dimers}, Kasteleyn
2496: considered imposing an energy function that favors one orientation of
2497: dimer over another (horizontal versus vertical).  The authors of
2498: \cite{gip} followed this lead, and showed how their methods also led
2499: to exact results for random domino tilings of the Aztec diamond when
2500: the distribution was skewed towards dominos of a particular
2501: orientation.  Here, we will state the results that follow from
2502: applying the methods of this paper to the case of biased tilings.
2503: 
2504: Let $p$ be strictly between $0$ and $1$.  For each $n$, there is a
2505: unique probability distribution on the tilings of the Aztec diamond of
2506: order $n$ (in fact, on any simply-connected region) such that given
2507: any tiling of all of the diamond except for a $2 \times 2$ block, the
2508: conditional probability that the $2 \times 2$ block will contain two
2509: horizontal dominos is $p$.  For more details on this distribution, see
2510: \cite{circle} or \cite{gip}.  We call this the Gibbs distribution with
2511: bias $p$.  (For more information on Gibbs distributions in general,
2512: see \cite{georgii}.)
2513: 
2514: The main difference between the biased distribution and the uniform
2515: distribution is the shape of the temperate zone.  We will see shortly
2516: that, in the biased case, its boundary is given by the ``arctic
2517: ellipse'' $\frac{x^2}{p}+\frac{y^2}{1-p}=1$ (in normalized
2518: coordinates).  It was conjectured in \cite{circle} that the analogue
2519: of the arctic circle theorem holds in the biased case.  Our methods
2520: prove that conjecture, as well as an arctangent formula that describes
2521: the behavior within the temperate zone.
2522: 
2523: We begin by defining the biased placement probabilities
2524: $\Pl_p(\ell,m;n)$ the same way we defined the ordinary placement
2525: probabilities (except, of course, that we use the biased
2526: distribution).  The biased creation rates are also defined analogously
2527: to the ordinary creation rates, by
2528: $$
2529: \Cr_p(\ell,m;n) = \frac{1}{p}(\Pl_p(\ell,m;n) - \Pl_p(\ell,m-1;n-1)).
2530: $$
2531: 
2532: The proofs depend on a biased version of \propref{gessel}, which is
2533: proved in \cite{gip}.  To state it, we will need a more general form
2534: of Krawtchouk polynomial.  Define $c_p(a,b;n)$ to be the coefficient
2535: of $z^a$ in $(1+(1-p)z/p)^{n-b}(1-z)^b.$ (See \cite[p.~151]{codes}.)
2536: We need the following result from \cite{gip}:
2537: 
2538: \begin{prop}
2539: \label{biasgessel}
2540: Let $0 < p < 1$, and set $a = (\ell+m+n)/2$ and $b=(\ell-m+n)/2$.  If
2541: $a,b \in \Z$, then
2542: $$
2543: {\Cr_p}(\ell,m;n+1) =
2544: c_p(a,b;n)c_p(b,a;n)p^n.
2545: $$
2546: Otherwise, ${\Cr_p}(\ell,m;n+1) = 0.$
2547: \end{prop}
2548: 
2549: Using this proposition, straightforward modifications to the proof of
2550: \propref{crest} prove the following generalization:
2551: 
2552: \begin{prop}
2553: \label{biascrest}
2554: Fix $\varepsilon > 0$.  If $\frac{\ell^2}{p} + \frac{m^2}{1-p} \le
2555: (1-\varepsilon)n^2$ and $\ell+m \equiv n \pmod{2}$, then
2556: $$
2557: \Cr_p(\ell,m;n+1) = \frac{2\cos^2
2558: \Phi_p(\ell,m;n)}{\pi\sqrt{(p-p^2)n^2-(1-p)\ell^2-pm^2}} +
2559: O_\varepsilon(n^{-2})
2560: $$
2561: for some function $\Phi_p(\ell,m;n)$, which can be determined
2562: explicitly.
2563: \end{prop}
2564: 
2565: Every result needed for the proof of \thmref{main} (such as the
2566: creation rate estimates outside the arctic ellipse) has a
2567: straightforward generalization to the biased case; in the interest of
2568: saving space, we will omit their statements.  The proofs are
2569: completely analogous to the proofs for the uniform distribution.  One
2570: arrives at the following biased counterpart to \thmref{main}:
2571: 
2572: \begin{thm}
2573: \label{biasmain}
2574: Let $0 < p < 1$, and let $U$ be an open set containing the points
2575: $(\pm p, 1-p)$.  If $(x,y)$ is the normalized location of a
2576: north-going domino space in the Aztec diamond of order $n$, and $(x,y)
2577: \not\in U$, then, as $n \rightarrow \infty$, the placement probability
2578: at $(x,y)$ for the Gibbs distribution with bias $p$ is within $o(1)$
2579: of ${\mathcal P}_p(x,y)$, where
2580: $$
2581: {\mathcal P}_p(x,y) = 
2582: \begin{cases}
2583: 0&\!\!\!\!\hbox{if $\frac{x^2}{p}+\frac{y^2}{1-p}\ge1$ and $y<1-p$,}\\
2584: 1&\!\!\!\!\hbox{if $\frac{x^2}{p}+\frac{y^2}{1-p}\ge1$ and $y>1-p$,
2585: and}\\ 
2586: \frac{1}{2}+\frac{1}{\pi}\tan^{-1}\!\!\left(\!\frac{y-(1-p)}
2587: {\sqrt{p-p^2-(1-p)x^2-py^2}}\!\right)\!
2588: &\!\!\!\!\hbox{if $\frac{x^2}{p}+\frac{y^2}{1-p}<1$}.\\
2589: \end{cases}
2590: $$
2591: The $o(1)$ error bound is uniform in $(x,y)$ (for $(x,y) \not\in U$).
2592: \end{thm}
2593: 
2594: Similarly, the south-going, east-going, and west-going placement
2595: probabilities near $(x,y)$ are approximated by ${\mathcal
2596: P}_p(-x,-y)$, ${\mathcal P}_{1-p}(-y,x)$, and ${\mathcal
2597: P}_{1-p}(y,-x)$, respectively. This follows from \thmref{biasmain} by
2598: rotational symmetry.
2599: 
2600: One can also prove biased versions of the robustness and variance
2601: results from
2602: subsections~\ref{ssec-robustness}~and~\ref{ssec-variance}.  (In fact,
2603: the proofs are practically identical to the proofs given in those
2604: subsections.)  Using them in combination with the same methods used in
2605: subsection~\ref{ssec-ACT}, we can prove a slightly strengthened
2606: version of the ``arctic ellipse conjecture'' from \cite{circle}:
2607: 
2608: \begin{thm}
2609: \label{aec}
2610: Let $0 < p < 1$, and $\varepsilon > 0$.  The probability that, in a
2611: random domino tiling with bias $p$ of an Aztec diamond of order $n$,
2612: the boundary of the polar regions is more than a distance
2613: $\varepsilon$ in normalized coordinates from the ellipse
2614: $\frac{x^2}{p}+\frac{y^2}{1-p} = 1$ is exponentially small in $n$.
2615: \end{thm}
2616: 
2617: \section{Speculations}
2618: \label{sec-speculations}
2619: 
2620: In this article we have focused primarily on one particular family of
2621: finite regions, namely, Aztec diamonds.  Here we will indicate what it
2622: might mean to have a theory that would apply to {\sl all}
2623: simply-connected finite regions, and how Aztec diamonds might play a
2624: role in the project of classifying the different possible local
2625: behaviors that random tilings of such regions can exhibit away from
2626: their boundaries.
2627: 
2628: The results of
2629: subsections~\ref{ssec-robustness}~and~\ref{ssec-variance}~tell us that
2630: for any large simply-connected region $R$ that can be tiled by
2631: dominos, height functions associated with random tilings of $R$ will
2632: cluster around their average.  We furthermore know that this average
2633: height function depends in a monotone way on the values of the height
2634: function on the boundary of $R$, and is stable under certain kinds of
2635: slight perturbations of the boundary of $R$.  However, what these
2636: theorems do not tell us is whether this dependence is robust under
2637: scaling as well.  \propref{heightformula} tells us that such
2638: robustness does in fact hold for Aztec diamonds: that is, when one
2639: normalizes two large Aztec diamonds, one finds that the normalized
2640: average height functions are very close to one another.  That this is
2641: true along the boundary is a triviality; that it is true in the
2642: interior is a much subtler property, known to us only as a consequence
2643: of \thmref{main}.
2644: 
2645: We conjecture that scaling-robustness of height functions is true in
2646: general.  That is, suppose $R_1,R_2,\dots$ are finite,
2647: simply-connected, domino-tileable regions that grow without bound,
2648: such that suitably rescaled copies of the $R_n$'s converge to some
2649: compact subset $R^*$ of the plane.  Moreover, suppose that the height
2650: functions associated with the boundaries of the $R_n$'s, when rescaled
2651: by the same respective amounts, converge to some function on the
2652: boundary of $R^*$.  Then we believe that the average height functions
2653: associated with the $R_n$'s, when rescaled, converge on the interior
2654: of $R^*$ as well as on the boundary to some function ${\mathcal H}$.
2655: If the boundary values behave reasonably (perhaps piecewise smoothness
2656: suffices), then ${\mathcal H}$ should be piecewise smooth (with
2657: reasonably shaped pieces).
2658: 
2659: Under this picture, we view ${\mathcal H}$ as the solution to a
2660: somewhat strange sort of Dirichlet problem.  We will have more to say
2661: about this analogy shortly, but first we must leave the issues of
2662: large-scale structure (embodied in the average height function) and
2663: discuss the small-scale structure of random tilings.
2664: 
2665: Consider simply-connected regions $R_1,R_2,\dots$ as above.  In each
2666: region $R_n$, choose a north-going domino space $\sigma_n$ with
2667: normalized location $(x_n,y_n)$ in $R^*$, so that $(x_n,y_n)
2668: \rightarrow (x^*,y^*)$ as $n \rightarrow \infty$, and suppose that the
2669: asymptotic renormalized height function ${\mathcal H}$ is
2670: differentiable at $(x^*,y^*)$.  Assume that $(x^*,y^*)$ is in the
2671: interior of $R^*$ and that ${\mathcal H}$ is ``non-extremal'' at
2672: $(x^*,y^*)$, in the sense that its tilt $(s,t) = (\frac{\partial
2673: {\mathcal H}}{\partial x},\frac{\partial {\mathcal H}}{\partial y})$
2674: satisfies $|s|+|t|<2$.  Then we conjecture that the placement
2675: probabilities at the chosen north-going domino spaces $\sigma_n$
2676: converge.  The arctangent formula tells us that the conjecture is in
2677: fact true for Aztec diamonds.
2678: 
2679: Note that if we were to replace each $\sigma_n$ by another north-going
2680: domino space $\sigma'_n$ obtained by shifting it by some fixed vector
2681: $(i,j)$ with $i+j$ even, we would get the same point $(x^*,y^*)$ in
2682: the normalized limit.  Hence, the preceding conjecture implies
2683: approximate local translation-invariance for the first-order
2684: statistics governing random tilings of large regions, provided one
2685: stays away from the boundary (and the tilt is non-extremal).
2686: 
2687: This corollary gives us a way to understand the importance of our
2688: hypothesis of non-extremality.  For instance, consider the region
2689: shown in Figure~\ref{fig-herringbone}; it has only one tiling, whose
2690: local statistics are in no sense governed by any of the statistics
2691: seen in Aztec diamonds.  Taking a suitable limit of such regions one
2692: gets a height function whose tilt $(s,t)$ satisfies $|s|+|t|=2$ and
2693: hence violates non-extremality.  Indeed, the statistics do not even
2694: exhibit local translation-invariance.  (Note also that for the Aztec
2695: diamond itself, ${\mathcal H}(\cdot,\cdot)$ is extremal at $(x^*,y^*)$
2696: if and only if the asymptotic entropy at normalized location
2697: $(x^*,y^*)$ is zero, which is the case if and only if the asymptotic
2698: density of $2\times 2$ blocks at normalized location $(x^*,y^*)$ is
2699: zero.)
2700: 
2701: \begin{figure}
2702: \begin{center}
2703: \PSbox{herringbone.ps hoffset=-10}{1.5in}{1.5in}
2704: \end{center}
2705: \caption{Herringbone pattern.}
2706: \label{fig-herringbone}
2707: \end{figure}
2708: 
2709: Having made a conjecture about convergence of first-order statistics,
2710: we naturally wonder about higher-order statistics as well.  We
2711: conjecture that in fact all finite-order statistics in the vicinity of
2712: the points $(x_n,y_n)$ stabilize as $n \rightarrow \infty$, yielding
2713: statistics that in some sense ``belong'' to the limit point
2714: $(x^*,y^*)$ (as long as the tilt at $(x^*,y^*)$ is non-extremal).
2715: Then, applying the translation-invariance remark made in the preceding
2716: paragraph, it follows that each $(x^*,y^*)$ determines a process whose
2717: values are domino tilings of the entire plane.  For instance, taking
2718: the $R_n$'s to be Aztec diamonds and the point $(x^*,y^*)$ to be the
2719: center of the normalized diamond, it is natural to conjecture that at
2720: the center of the Aztec diamond of order $n$, the local finite-order
2721: statistics converge to those of the maximal entropy process mentioned
2722: in the Introduction.  (This special case was conjectured in
2723: \cite{circle}.)  Letting $(x^*,y^*)$ vary inside the rescaled
2724: temperate zone, we would get a two-parameter family of tiling-valued
2725: processes; they would all be distinct from one another because they
2726: would have distinct first-order statistics.  The maximal entropy
2727: process would be unique among these processes not only in having the
2728: highest entropy but also in being invariant under the full group of
2729: lattice translations, rather than merely the color-preserving subgroup
2730: of index 2.
2731: 
2732: It can be shown rigorously that such processes, if they exist, have a
2733: combinatorial analogue of the ``Gibbs property'' studied in
2734: equilibrium statistical mechanics; that is, given a tiling of a
2735: cofinite subset of the plane whose finite complement is tileable, if
2736: one conditions the random process on that particular tiling, then the
2737: conditional distribution on tilings of the entire plane is uniform.
2738: 
2739: Here we leave aside caution and put forward some conjectures about
2740: what sort of shape the ultimate theory we are striving towards will
2741: take.  These surmises might be false, but we believe they are the
2742: natural avenues to pursue in further investigations of the theory.
2743: 
2744: In the first place, we conjecture that the tiling-valued processes
2745: associated with the points $(x^*,y^*)$ will turn out to be ergodic, or
2746: indecomposable, in the usual sense of the theory of dynamical systems.
2747: It is not hard to use the ergodic theorem for $\Z^2$-actions (see
2748: \cite{ergodic}) to show that every ergodic, translation-invariant
2749: (under color-preserving translations), tiling-valued random process
2750: determines placement probabilities $p_n$, $p_s$, $p_w$, and $p_e$ and
2751: thence determines a tilt $(s,t)=(2(p_w-p_e),2(p_n-p_s))$.  We predict
2752: that in those cases where the tilt is non-extremal (i.e., $|s|+|t|$ is
2753: strictly less than 2), there is in fact a unique ergodic Gibbs measure
2754: with tilt $(s,t)$.  If this were true, it would have many nice
2755: consequences; for instance, the four numbers $p_n,p_s,p_e,p_w$ would
2756: all be determined by the pair $(s,t)$, and thus would exhibit only two
2757: degrees of freedom, despite the fact that the only obvious constraint
2758: governing them is $p_n+p_s+p_e+p_w=1$.  A further nice property is
2759: that the temperate zones of Aztec diamonds would be universal in the
2760: sense that they would manifest, in the limit, all possible forms of
2761: non-extremal local behavior that random tilings of large
2762: simply-connected regions can manifest away from boundaries.  This
2763: universality is not peculiar to Aztec diamonds, but instead arises
2764: {}from the fact, proved in subsection~\ref{ssec-heterogeneity}, that
2765: Aztec diamonds exhibit all possible non-extremal tilts.
2766: 
2767: An especially nice benefit of the preceding conjecture is that it
2768: would open the door to a variational approach to the problem of
2769: finding the average height function $\mathcal H$ on $R^*$ given only
2770: its values on the boundary of $R^*$.  Given any candidate for
2771: $\mathcal H$, define $N_n$ as the number of domino tilings of $R_n$
2772: whose normalized height functions stay close to $\mathcal H$.  It does
2773: not seem too far-fetched to hope that the logarithm of $N_n$, when
2774: divided by the area of $R_n$, converges to an integral over $R^*$, in
2775: which the integrand is the entropy associated with the unique ergodic
2776: Gibbs process with tilt $(\frac{\partial {\mathcal H}}{\partial x},
2777: \frac{\partial {\mathcal H}}{\partial y})$.  Since finding the average
2778: height function on $R_n$ corresponds in some sense to maximizing
2779: $N_n$, we would hope that finding the asymptotic normalized height
2780: function on $R^*$ corresponds to maximizing this integral.  It might
2781: not always be possible to solve the associated calculus of variations
2782: problem explicitly, but such a theorem would be a major advance
2783: towards a complete understanding of how the presence of boundary
2784: conditions can affect the behavior of a domino tiling in the interior
2785: of a region.
2786: 
2787: The preceding idea has in fact been used by physicists, in the context
2788: of crystals; see for example \cite[pp.~3562--3563]{nhb}.  There, it is
2789: claimed that the shape of a crystal surface is determined by
2790: minimizing the total surface free energy, which is obtained by
2791: integrating a local contribution (the surface free energy density)
2792: depending only on the gradient of the surface.  This is believable
2793: physically, but in any particular lattice model it seems difficult to
2794: establish rigorously; it is not even clear on purely mathematical
2795: grounds why there should exist a surface free energy density depending
2796: only on the gradient.  The analogous statement in random tiling theory
2797: is the existence of a local entropy depending only on the tilt of the
2798: height function, but it is conceivable (although we consider it
2799: unlikely) that the local entropy might not be determined by the local
2800: asymptotic behavior of the normalized height function.  The only
2801: approach that we know of that might lead to a rigorous proof (or even
2802: a heuristic argument) is to prove the conjectures above about local
2803: statistics and Gibbs measures.
2804: 
2805: \section*{Acknowledgements}
2806: 
2807: We thank Robin Pemantle for suggesting the idea behind
2808: \lemref{monotone}.  Thanks also to David Feldman for providing helpful
2809: comments on the manuscript, to Sameera Iyengar for writing the program
2810: that produced the random tiling shown in Figure~\ref{fig-aztec}, to
2811: M.~Josie Ammer and Dan Ionescu for writing the first programs to
2812: compute placement probabilities for Aztec diamonds, to Pramod Achar,
2813: Federico Ardila, Dan Ionescu, and Ben Raphael for helping to write the
2814: programs that produced the random tilings shown in
2815: Figures~\ref{fig-fake1}, \ref{fig-fake2}, and \ref{fig-real1}, and to
2816: David Wilson for helping to create the figures.
2817: 
2818: \begin{thebibliography}{EKLP}
2819: 
2820: \bibitem[AS]{method}  N.\ Alon and J.\ Spencer,
2821: {\sl The probabilistic method\/},
2822: Wiley, New York, 1992.
2823: 
2824: \bibitem[vB]{bcsos} H.~van Beijeren, {\sl Exactly solvable model for
2825: the roughening transition of a crystal surface\/}, Phys.\ Rev.\ Lett.\
2826: {\bf 38} (1977), 993--996.
2827: 
2828: \bibitem[dB]{asymp} N.~G.\ de Bruijn, 
2829: {\sl Asymptotic methods in analysis\/}, 
2830: Dover Publications, Inc., New York, 1981.
2831: 
2832: \bibitem[BH]{bh} H.~W.~J.\ Bl\"ote and H.~J.\ Hilhorst,
2833: {\sl Roughening transitions and the zero-temperature triangular Ising
2834: antiferromagnet\/},  J.\ Phys.\ A {\bf 15} (1982), L631--L637.
2835: 
2836: \bibitem[BP]{local}   R.\ Burton and R.\ Pemantle,
2837: {\sl Local characteristics, entropy and limit theorems for spanning
2838: trees and domino tilings via transfer-impedances\/},
2839: Ann.\ Probab. {\bf 21} (1993), 1329--1371.
2840: 
2841: \bibitem[D]{durrett}  R.\ Durrett,
2842: {\sl Probability: theory and examples\/},
2843: Wadsworth and Brooks/Cole, Pacific Grove, California, 1991.
2844: 
2845: \bibitem[EKLP]{alternating} 
2846: N.\ Elkies, G.\ Kuperberg, M.\ Larsen, and J.\ Propp, 
2847: {\sl Alternating sign matrices and domino tilings\/},
2848: J.\ Algebraic Combin.\ {\bf 1} (1992),
2849: 111--132 and 219--234.
2850: 
2851: \bibitem[E]{elser} V.\ Elser, {\sl Solution of the dimer problem on
2852: a hexagonal lattice with boundary\/}, J.\ Phys.\ A {\bf 17} (1984),
2853: 1509--1513. 
2854: 
2855: \bibitem[G]{georgii} H.-O.\ Georgii, {\sl Gibbs measures and
2856: phase transitions\/}, De Gruyter, New York, 1988.
2857: 
2858: \bibitem[GIP]{gip} I.\ Gessel, A.\ Ionescu, and J.\ Propp, 
2859: {\sl Counting constrained domino tilings of Aztec diamonds\/},
2860: preprint, 1996.
2861: 
2862: \bibitem[GK]{expsums} S.~W.\ Graham and G.\ Kolesnik, {\sl Van der
2863: Corput's method of exponential sums\/}, Cambridge University Press,
2864: Cambridge, 1991.
2865: 
2866: \bibitem[GG]{gg} D.\ Grensing and G.\ Grensing, {\sl Boundary effects
2867: in the dimer problem on a non-Bravais lattice\/}, J.\ Math.\
2868: Phys.\ {\bf 24} (1983), 620--630.
2869: 
2870: \bibitem[JPS]{circle} W.\ Jockusch, J.\ Propp, and P.\ Shor,
2871: {\sl Random domino tilings and the arctic circle theorem\/},
2872: preprint, 1995.
2873: 
2874: \bibitem[Ka]{dimers} P.~W.\ Kasteleyn,
2875: {\sl The statistics of dimers on a lattice, I. 
2876: The number of dimer arrangements on a quadratic lattice\/},
2877: Physica {\bf 27} (1961), 1209--1225.
2878: 
2879: \bibitem[Kr]{ergodic} U.\ Krengel,
2880: {\sl Ergodic theorems\/},
2881: W.\ de Gruyter, New York, 1985.
2882: 
2883: \bibitem[L]{levitov} L.~S.\ Levitov, {\sl Equivalence of the dimer
2884: resonating-valence-bond problem to the quantum roughening problem\/},
2885: Phys.\ Rev.\ Lett.\ {\bf 64} (1990), 92--94.
2886: 
2887: \bibitem[MS]{codes} F.~J.\ MacWilliams and N.~J.~A.\ Sloane,
2888: {\sl The theory of error-correcting codes\/}, North-Holland Publishing
2889: Company, Amsterdam, 1978.
2890: 
2891: \bibitem[NHB]{nhb} B.\ Nienhuis, H.~J.\ Hilhorst, and H.~W.~J.\
2892: Bl\"ote, {\sl Triangular {\sc SOS} models and cubic-crystal shapes\/},
2893: J.\ Phys.\ A {\bf 17} (1984), 3559--3581.
2894: 
2895: \bibitem[PW]{exact} J.~G.\ Propp and D.~B.\ Wilson,
2896: {\sl Exact sampling with coupled Markov chains and applications to
2897: statistical mechanics\/}, to appear in Random Structures and
2898: Algorithms.
2899: 
2900: \bibitem[R]{rauch} J.\ Rauch, {\sl Partial differential
2901: equations\/}, Springer-Verlag, New York, 1991.
2902: 
2903: \bibitem[STCR]{spaces}
2904: N.\ Saldanha, C.\ Tomei, M.\ Casarin, Jr., and D.\  Romualdo, 
2905: {\sl Spaces of domino tilings\/}, Discrete Comput.\
2906: Geom.\ {\bf 14} (1995), 207--233.
2907: 
2908: \bibitem[SZ]{remark} H.\ Sachs and H.\ Zernitz,
2909: {\sl Remark on the dimer problem\/},
2910: Discrete Appl.\ Math.\ {\bf 51} (1994), 171--179.
2911: 
2912: \bibitem[T]{conway} W.~P.\ Thurston,
2913: {\sl Conway's tiling groups\/},
2914: Amer.\ Math.\ Monthly {\bf 97} (1990), 757--773.
2915: 
2916: \end{thebibliography}
2917: 
2918: \end{document}
2919: 
2920: