1: \documentclass[12pt]{article}
2: \usepackage{amsmath,amssymb,graphicx,latexsym}
3: \def\trace{\rm trace}\newcommand{\qed}{\hfill\vrule height6pt width6pt
4: depth0pt \medskip}
5: \newcommand{\pr}{\mathbb{P}}
6: \newcommand{\arct}{\tan^{-1}}
7: \newcommand{\arcs}{\sin^{-1}}
8: \newcommand{\arcc}{\cos^{-1}}
9: \newcommand{\sgn}{\textrm{sgn}\,}
10: \newcommand{\sq}{\sqrt{2}/2}
11: \newcommand{\tsq}{\theta\sqrt{2}/2}
12: \newcommand{\bo}{\beta_1}
13: \newcommand{\bt}{\beta_2}
14: \newcommand{\gamo}{\gamma_1}
15: \newcommand{\gamt}{\gamma_2}
16: \newcommand{\nequiv}{\equiv\!\!\!\!\!\!/\ }
17: \title{Limit shapes for random square Young tableaux and plane
18: partitions}
19: \author{Dan Romik, Boris Pittel}
20: \linespread{1.2}
21:
22: \begin{document}
23: %\setlength{\parindent}{0pt}
24: %\maketitle
25:
26: \begin{center} \textsc{\large{Limit shapes for random square Young
27: tableaux and plane partitions}} \\\ \\
28: \footnotesize
29: \begin{tabular}{cc}
30: \textsc{\large{Boris Pittel}}
31: \footnote{Supported in part by the NSF Grant DMS-0104104.}
32: & \textsc{\large{Dan Romik}} \\
33: Department of Mathematics & Department of Mathematics \\
34: Ohio State University & The Weizmann Institute of Science \\
35: Columbus, Ohio 43210 & Rehovot 76100, Israel \\
36: email: \tt{bgp@math.ohio-state.edu}
37: & email: \tt{romik@wisdom.weizmann.ac.il}
38: \end{tabular}
39: \end{center}
40:
41:
42: \footnotesize
43: \paragraph{Abstract.}Our main result is a limit shape theorem for the
44: two-dimensional surface defined by a uniform random $n\times n$ square
45: Young tableau. The analysis leads to a calculus of variations
46: minimization problem that resembles the minimization problems studied
47: by Logan-Shepp, Vershik-Kerov, and Cohn-Larsen-Propp. Our solution
48: involves methods from the theory of singular integral equations, and
49: sheds light on the somewhat mysterious derivations in these works. An
50: extension to rectangular diagrams, using the same ideas but involving
51: some nontrivial computations, is also given.
52:
53: We give several applications of the main result. First, we show that
54: the location of a particular entry in the tableau is in the limit
55: governed by a semicircle distribution.
56:
57: Next, we derive a result on the length of the longest increasing
58: subsequence in segments of a \emph{minimal Erd\"os-Szekeres
59: permutation}, namely a permutation of the numbers $1,2,\ldots,n^2$ whose
60: longest monotone subsequence is of length $n$ (and hence minimal by
61: the Erd\"os-Szekeres theorem).
62:
63: Finally, we prove a limit shape theorem for the surface defined by a
64: random plane partition of a very large integer over a large square
65: (and more generally rectangular) diagram.
66:
67: \vspace{40.0 pt}
68: \noindent
69: {\bf 2000 Mathematics Subject Classifications:} Primary 60C05;
70: Secondary 05E10, 60F10.
71:
72: \vspace{45.0 pt}
73: \noindent
74: \begin{tabular}{c} \qquad\qquad\qquad
75: \qquad\qquad\qquad\qquad\qquad\qquad \\\hline \end{tabular}
76:
77: $^1$Supported in part by the NSF Grant DMS-0104104.
78:
79:
80: \newpage
81: \tableofcontents
82:
83:
84: \newpage
85: \section{Introduction}
86:
87: \subsection{Random square Young tableaux}
88:
89: In this paper, we study the large-scale asymptotic behavior of uniform
90: random Young tableaux chosen from the set of tableaux of square
91: shape. Recall that a \emph{Young diagram} is a graphical
92: representation of a partition $\lambda:
93: \lambda(1)\ge\lambda(2)\ge\ldots\ge\lambda(k)$ of $n=\sum \lambda_i$ as an
94: array of \emph{cells}, where row $i$ has $\lambda_i$ cells. For a
95: Young diagram $\lambda$ (we will often identify a partition with its
96: Young diagram), a \emph{Young tableau} of shape $\lambda$ is a filling
97: of the cells of $\lambda$ with the numbers $1,2,\ldots,n$ such that the
98: numbers along every row and column are increasing.
99:
100: A square Young tableau is a Young tableau whose shape is an $n\times
101: n$ square Young diagram. The number of such tableaux is
102: known by the hook formula of Frame-Thrall-Robinson (see
103: \eqref{eq:hook} below) to be
104: $$ \frac{(n^2)!}{[1\cdot (2n-1)][2\cdot(2n-2)]^2[3\cdot(2n-3)]^3
105: \ldots [(n-1)(n+1)]^{n-1} \ n^n }. $$ A square tableau
106: $T=(t_{i,j})_{i,j=1}^n$ can be depicted geometrically as a
107: three-dimensional stack of cubes over the two-dimensional square
108: $[0,n]\times[0,n]$, where $t_{i,j}$ cubes are stacked over the square
109: $[i-1,i]\times[j-1,j]\times\{0\}$. Alternatively, the function
110: $(i,j)\to t_{i,j}$ can be thought of as the graph of the
111: (non-continuous) surface of the upper envelope of this stack. By
112: rescaling the $n\times n$ square to a square of unit sides, and
113: rescaling the heights of the columns of cubes so that they are all
114: between 0 and 1, one may consider the family of square tableaux as
115: $n\to\infty$. This raises the natural question, whether the shape of
116: the stack for a \emph{random} $n\times n$ square tableau exhibits some
117: asymptotic behavior as $n\to\infty$. The answer is given by the
118: following theorem, and is illustrated in Figure 1.
119:
120: \paragraph{Theorem 1.} Let ${\cal T}_n$ be the set of $n\times n$
121: square Young tableaux, and let $\pr_n$ be the uniform probability
122: measure on ${\cal T}_n$. Then for the function
123: $L:[0,1]\times[0,1]\to[0,1]$ defined below, we have:
124:
125: (i) \emph{Uniform convergence to the limit shape:} for all $\epsilon>0$,
126: $$ \pr_n\left( T\in{\cal T}_n : \max_{1\le i,j\le n}
127: \left|\frac{1}{n^2}t_{i,j} -
128: L\left(\frac{i}{n},\frac{j}{n}\right)\right| > \epsilon \right)
129: \xrightarrow[n\to\infty]{} 0.$$
130:
131: (ii) \emph{Rate of convergence in the interior of the square:} for all
132: $\epsilon>0$,
133: $$ \pr_n\bigg( T\in{\cal T}_n : \max_{ \tiny{
134: \begin{array}{c} 1\le i,j\le n \\
135: \min(i j,(n-i)(n-j)) > n^{3/2+\epsilon} \end{array}}}
136: \left|\frac{1}{n^2}t_{i,j} - L\left(\frac{i}{n},\frac{j}{n}\right)\right| >
137: \frac{1}{n^{(1-\epsilon)/2}} \bigg)
138: \xrightarrow[n\to\infty]{} 0. $$
139:
140: \newpage
141:
142: \begin{figure}[h!]
143: %\begin{center}
144: \hspace{-50.0pt}
145: \hbox{$
146: \begin{array}{c}
147: \includegraphics{tableau3d.eps} \\
148: \textrm{\footnotesize{(a) 3D plot of simulated tableau}}
149: \end{array}$ $
150: \begin{array}{c}
151: \includegraphics{limit3d.eps} \\
152: \textrm{\footnotesize{(b) The limit surface $L(x,y)$}}
153: \end{array}$
154: }
155: \medskip
156: \hspace{-50.0pt}
157: \hbox{$
158: \begin{array}{c}
159: \includegraphics{tableaucontour.eps} \\
160: \textrm{\footnotesize{(c) Contour plot of simulated tableau}}
161: \end{array}$ $
162: \begin{array}{c}
163: \includegraphics{limitcontour.eps} \\
164: \textrm{\footnotesize{(d) Contour plot of $L$}}
165: \end{array}$
166: }
167: \caption{A simulated $50\times 50$ random tableau and the limit surface}
168: %\end{center}
169: \end{figure}
170:
171: \paragraph{Definition of $L$.} We call the function $L$ the
172: \emph{limit surface of square Young tableaux}. It is defined by the
173: implicit equation
174: $$ x+y =\frac{2}{\pi}(x-y)\arct\left(
175: \frac{ (1-2L(x,y))(x-y)}{\sqrt{4L(x,y)(1-L(x,y)) -
176: (x-y)^2}} \right) $$
177: $$ + \frac{2}{\pi}\arct \left(\frac{\sqrt{4L(x,y)(1-L(x,y))-
178: (x-y)^2}}{1-2L(x,y)}\right) $$
179: for $0 \le y \le 1-x \le 1$, together with the reflection property
180: $$ L(x,y) = 1- L(1-x,1-y) $$
181: (where $\arct$ is the arctangent function).
182: It is more natural to describe $L$ in terms of its level
183: curves $\{L(x,y)=\alpha\}$. First, introduce the \emph{rotated coordinate
184: system}
185: \begin{equation}\label{eq:rotated}
186: u=\frac{x-y}{\sqrt{2}},\qquad v=\frac{x+y}{\sqrt{2}}.
187: \end{equation}
188: In the $u-v$ plane, the
189: square $[0,1]\times[0,1]$ transforms into the rotated square
190: $$ \Diamond = \{ (u,v)\in \mathbb{R}^2 : |u|\le \sqrt{2}/2,
191: |u|\le v \le \sqrt{2}-|u| \}. $$
192: Now define the one-parameter family of functions
193: $(g_\alpha)_{0\le\alpha\le 1}$ given by
194: $$g_\alpha : [-\sqrt{2\alpha(1-\alpha)},\sqrt{2\alpha(1-\alpha)}]\to
195: \mathbb{R}, $$
196: \begin{multline}\label{eq:minimizers} {\scriptsize
197: g_\alpha(u) = \left\{ \begin{array}{ll}
198: \ \ \frac{2}{\pi}
199: u\arct\left(\frac{(1-2\alpha)u}{\sqrt{2\alpha(1-\alpha)-u^2}}
200: \right) +
201: \frac{\sqrt{2}}{\pi}
202: \arct
203: \left(\frac{\sqrt{2(2\alpha(1-\alpha)-u^2)}}{1-2\alpha}\right)
204: & 0 \le \alpha < \frac{1}{2}, \\
205: -\frac{2}{\pi}
206: u\arct\left(\frac{(1-2\alpha)u}{\sqrt{2\alpha(1-\alpha)-u^2}}
207: \right) -
208: \frac{\sqrt{2}}{\pi}
209: \arct
210: \left(\frac{\sqrt{2(2\alpha(1-\alpha)-u^2)}}{1-2\alpha}\right) +\sqrt{2}
211: & \frac{1}{2} < \alpha \le 1, \\
212: \frac{\sqrt{2}}{2} & \alpha=\frac{1}{2}.
213: \end{array}\right.}
214: \end{multline}
215: Then in the rotated coordinate system, the surface $\bar{L}(u,v) =
216: L(x(u,v),y(u,v))$ can be described as the surface whose level curves
217: $\{\bar{L}(u,v)=\alpha\}$ are exactly the curves
218: $\{v=g_\alpha(u)\}$. That is,
219: $$ \{ (u,v)\in\Diamond : \bar{L}(u,v) = \alpha \} =
220: \{ (u,v)\in\Diamond : |u|\le \sqrt{2\alpha(1-\alpha)},
221: v=g_\alpha(u) \}. $$
222: This is illustrated in Figure 2. It is straightforward to check that
223: the curves $v=g_\alpha(u)$ do not intersect, and so define a surface
224: \footnote{See equation \eqref{eq:partialdiff} in section 3.4.}.
225:
226: \begin{figure}
227: \begin{center}
228: \includegraphics{levelcurves.eps}
229: \caption{The curves $v=g_\alpha(u)$ for
230: $\alpha = 0.05,\ 0.1,\ 0.15,\ 0.2,\ \ldots\ ,\ 0.5$}
231: \end{center}
232: \end{figure}
233:
234: Note some special values of $L(x,y)$ which can be computed explicitly:
235: \begin{eqnarray*} L(t,0) &=& L(0,t)\ \, =\ \, \frac{1-\sqrt{1-t^2}}{2}, \\
236: L(t,1) &=& L(1,t)\ \,=\ \,\frac{1+\sqrt{2t-t^2}}{2}, \\
237: L(t,t) &=& \frac{1-\cos(\pi t)}{2}.
238: \end{eqnarray*}
239:
240: \bigskip
241: The approach in proving Theorem 1 is the variational approach. Namely,
242: we identify the large-deviation rate functional of the level curves of
243: the random surface defined by the tableau, then analyze the functional
244: and find its minimizers. This will give Theorem 1(ii), with the rate
245: of convergence following from classical norm estimates for some
246: integral operators. The treatment of the boundary of the square,
247: required for Theorem 1(i), turns out to be more delicate, and will
248: require special arguments.
249:
250: \subsection{Location of particular entries}
251:
252: Theorem 1 identifies the approximate value of the entry of a typical square
253: tableau in a given location in the square. A dual outlook is to ask
254: where a given value $k$ will appear in the square tableau, since all
255: the values between 1 and $n^2$ appear exactly once. These questions
256: are almost equivalent. Indeed, if $k$ is approximately $\alpha\cdot
257: n^2$, then Theorem 1 predicts that with high probability the entry $k$
258: will appear in the vicinity of the level curve $\{L(x,y) = \alpha\}$
259: (the fact that this actually follows from Theorem 1 is a simple
260: consequence of the monotonicity property of the tableau along rows and
261: columns). However, one may ask a more detailed question about the
262: limiting distribution of the location of the entry $k$ on the level
263: curve. It turns out that its $u$-coordinate has approximately the semicircle
264: distribution. This is made precise in the following theorem.
265:
266: \paragraph{Theorem 2.} For a tableau $T\in{\cal T}_n$ and $1\le k\le
267: n^2$, denote by $(i(T,k),j(T,k))$ the location of the entry $k$ in
268: $T$, and denote $X(T,k)=i(T,k)/n,\ Y(T,k)=j(T,k)/n$. Let $0<\alpha<1$,
269: let $k_n$ be a sequence of integers such that $k_n/n^2
270: \xrightarrow[n\to\infty]{} \alpha$, and for each $n$ let $T_n$ be a
271: uniform random tableau in ${\cal T}_n$. Then as $n\to\infty$, the
272: random vector $(X(T_n,k_n),Y(T_n,k_n))$ converges in distribution to
273: the random vector
274: $$ (X_\alpha, Y_\alpha) := \left( \frac{V_\alpha+U_\alpha}{2},
275: \frac{V_\alpha-U_\alpha}{2} \right), $$
276: where $U_\alpha$ is a random variable with density function
277: \begin{equation}\label{eq:semicircle}
278: f_{U_\alpha}(u) = \frac{\sqrt{2\alpha(1-\alpha)-u^2}}{\pi \alpha(1-\alpha)}
279: \mathbf{1}_{[-\sqrt{2\alpha(1-\alpha)},
280: \sqrt{2\alpha(1-\alpha)}]}(u)
281: \end{equation}
282: and $V_\alpha = g_\alpha(U_\alpha)$.
283:
284: \bigskip
285: Theorem 2 is one of several aspects of our work which shows a deep
286: connection to the work of Logan-Shepp and Vershik-Kerov on the limit
287: shape of Plancherel-random partitions - see section 8 for discussion.
288:
289: \subsection{Minimal Erd\"os-Szekeres permutations} The famous
290: Erd\"os-Szekeres theorem states that a permutation of $1,2,\ldots,n^2$
291: must have either an increasing subsequence of length $n$ or a
292: decreasing subsequence of length $n$. This can be proved using the
293: pigeon-hole principle, but also follows from the RSK correspondence
294: using the observation that a Young diagram of area $n^2$ must have
295: either width or height at least $n$.
296:
297: For the width and height of a Young diagram of area $n^2$ to be
298: \emph{exactly} $n$, the diagram must be a square. From the RSK
299: correspondence it thus follows that to each permutation of
300: $1,2,\ldots,n^2$ whose longest increasing subsequence and longest
301: decreasing subsequence have length exactly $n$, there correspond a
302: pair of square $n\times n$ Young tableaux. Such a permutation has the
303: minimal possible length of a longest \emph{monotone} subsequence, and
it seems appropriate to term such permutations \emph{minimal
Erd\"os-Szekeres permutations} (we are not aware of any previous
references to these permutations, aside from a brief mention in
\cite{knuth}, exercise 5.1.4.9).
As an application of our limit shape result, we will prove the
following result on the length of the longest increasing subsequence
when just an initial segment of a random minimal Erd\"os-Szekeres
permutation is read.
\paragraph{Theorem 3.} For each $n$, let $\pi_n$ be a uniform random minimal
Erd\"os-Szekeres permutation of $1,2,\ldots,n^2$. For $1\le k\le n^2$, Let
$l_{n,k}$ be the length of the longest increasing subsequence in the
sequence $\pi_n(1), \pi_n(2),\ldots,\pi_n(k)$. Denote $\alpha=k/n^2$, and
$\alpha_0=n^{-2/3+\epsilon}$. Then for any $\epsilon>0$, and
$\omega(n)\to\infty$ however slowly,
$$ \max_{\alpha_0\le k/n^2\le 1/2}
\pr( |l_{n,k} -
2\sqrt{\alpha(1-\alpha)}n|> \alpha_0^{1/2}\omega(n)n) \xrightarrow[n\to\infty]{} 0.
$$
Thus the random fluctuations of $l_{n,k}$ around $2\sqrt{\alpha(1-\alpha)}n$ are not
likely to be of order substantially larger than $n^{2/3}$.
\subsection{A limit surface for random square plane partitions}
Another probability model which was studied in the context of limit
shapes, is that of random plane partitions. If $\lambda$ is a Young
diagram, \emph{a plane partition of $n$ of shape $\lambda$} is an
array of positive integers $(p_{i,j})_{(i,j)\in\lambda}$ indexed by
the cells of $\lambda$, that sum to $n$ and which are weakly
decreasing along rows and columns, i.e., satisfy
$$ p_{i,j} \ge p_{i,j+1}, \qquad p_{i,j} \ge p_{i+1,j}. $$
Cerf and Kenyon \cite{cerfkenyon} proved a limit shape result for
random unrestricted plane partitions of an integer $n$. Cohn, Larsen
and Propp \cite{cohnlarspropp} proved a limit shape result for random plane
partitions whose three-dimensional graph is bounded inside a large box
of given relative proportions.
%A recent work related to non-uniform
%plane partitions is that of Kenyon, Okounkov and Sheffield
%\cite{kenyonokounkovsheff}.
We apply Theorem 1 to prove a limit shape result for random plane
partitions of $m$ defined over an $n\times n$ square Young diagram,
304: when $m$ is much greater than $n^6$. This can be related to square
305: Young tableaux by the observation that when a plane partition does not
306: contain repeated entries (which in the asymptotic regime described
307: above happens with high probability), the order structure on the
308: entries of the plane partition is a Young tableau. The precise result is
309: the following.
310:
311: \paragraph{Theorem 4.} For integers $n,m>0$, let ${\cal P}_{n,m}$ be the set of
312: plane partitions of $m$ of $n\times n$ square shape, and let
313: $\pr_{n,m}$ be the uniform probability measure on ${\cal
314: P}_{n,m}$. If $\pi=(p_{i,j})_{i,j=1}^n$ is an element of ${\cal
315: P}_{n,m}$, let its \emph{rescaled surface graph} be the function
316: $ \tilde{S}_\pi:[0,1)\times[0,1)\to[0,\infty)$ defined by
317: $$ \tilde{S}_\pi(x,y) = \frac{n^2}{m}p_{\lfloor n x\rfloor+1, \lfloor
318: n y\rfloor+1}. $$
319:
320: Suppose $m$ and $n$ are sequences of integers that tend to infinity in
321: such a way that $m/n^6\to\infty$. Then for all
322: $\epsilon>0,\ x,y\in[0,1)$ we have
323: $$ \pr_{n,m}( \pi\in {\cal P}_{n,m} : |\tilde{S}_\pi(x,y)-M(x,y)|>
324: \epsilon ) \xrightarrow[\qquad]{} 0, $$
325: where $M:[0,1]\times[0,1]\to[0,\infty)$ is given by
326: $$ M(x,y) = -\log(L(x,y)). $$
327:
328: \medskip
329: Theorem 4 may be related to a limiting case $\gamma\to\infty$ in the
330: limit shape result of Cohn-Larsen-Propp \cite{cohnlarspropp}. We have
331: not attempted to check this.
332:
333: \subsection{Random rectangular Young tableaux and plane partitions}
334:
335: The methods which we will use to prove Theorems 1, 2, and 4 work
336: equally well for rectangular Young tableaux and plane partitions, in
337: the limit when the size of the rectangle grows and its relative
338: proportions tend to a limiting value $\theta>0$. For each possible value
339: $\theta$ of the ratio between the sides of the rectangle, there is a
340: limiting surface $L_\theta$ for random rectangular Young tableaux, and
341: a limiting surface $M_\theta$ for random rectangular plane
342: partitions.
343: %We have chosen to focus in this paper on square tableaux
344: %and plane partitions, because of the elegance of the formulas which
345: %appear in the analysis, because of the application to minimal
346: %Erd\"os-Szekeres permutations, and because of the special properties
347: %which makes this problem analgous to the problem of the limit shape of
348: %Plancherel-random partitions. We now state the generalization of our
349: %results to a rectangular shape. The proofs involve exactly the same
350: %ideas as for the square case, and simply require some additional
351: %computations, which we include at the end of the paper.
352: Analogously to the square tableaux, the rectangular $n_1\times n_2$
353: tableaux can be viewed as the result of applying the RSK algorithm to
354: a permutation of $\{1,\ldots,n_1 n_2\}$ with the property that the
355: lengths of the longest increasing and the longest decreasing
356: subsequences are exactly equal $n_1$ and $n_2$ (by the
357: Erd\"os-Szekeres theorem, the two lengths cannot be simultaneously
358: below $n_1$ and $n_2$, respectively). The proofs, which we include at
359: the end of the paper, require some nontrivial modifications, but the
360: final results are unexpectedly as elegant as for the square case.
361:
362: Let $\theta>0$. We may assume that $\theta \le 1$, otherwise exchange
363: the two sides of the rectangle. Define
364: $L_\theta:[0,1]\times[0,\theta]\to[0,1]$, the \emph{limit surface of
365: rectangular tableaux with side ratio $\theta$}, as follows. For each
366: $0<\alpha<1$, the $\alpha$-level curve $\{(x,y):
367: L_\theta(x,y)=\alpha\}$ is given in rotated $u-v$ coordinates by
368: $$ \{ (u,h_{\theta,\alpha}(u)) : -\beta_1 \le u \le
369: \beta_2 \}, $$ where
370: %
371: %
372: %
373: %\begin{eqnarray*}
374: %\beta_1 &=& \sqrt{2\theta\alpha(1-\alpha)} -
375: %\alpha(1-\theta)\sqrt{2}/2, \\ \beta_2 &=&
376: %\sqrt{2\theta\alpha(1-\alpha)} + \alpha(1-\theta)\sqrt{2}/2, \\
377: %h_{\theta,\alpha}(u) &=& \beta_1
378: %+ \frac{2\sqrt{2\theta\alpha(1-\alpha)}}{\pi}
379: %\bigg[ -(\xi+\gamo) \arct
380: %\sqrt{\frac{(1-\xi)(\gamo-1)}{(1+\xi)(\gamo+1)}} \\ & & + (\xi-\gamt) \arct
381: %\sqrt{\frac{(1+\xi)(\gamt-1)}{(1-\xi)(\gamt+1)}} \\ & & +
382: %\frac{1}{2}\left(\arcs \xi+\frac{\pi}{2}\right)
383: %\left(\sqrt{\gamt^2-1}-\sqrt{\gamo^2-1}\right) +
384: %\frac{\pi}{2}(\gamo-1) \bigg], \qquad\ \ \ 0< \alpha \le
385: %\frac{\theta}{\theta+1}, \\
386: %h_{\theta,\alpha}(u) &=& \theta\sqrt{2}-\beta_1
387: %+ \frac{2\sqrt{2\theta\alpha(1-\alpha)}}{\pi}
388: %\bigg[ (\xi+\gamo) \arct
389: %\sqrt{\frac{(1-\xi)(\gamo-1)}{(1+\xi)(\gamo+1)}} \\ & & + (\xi-\gamt) \arct
390: %\sqrt{\frac{(1+\xi)(\gamt-1)}{(1-\xi)(\gamt+1)}} \\ & & +
391: %\frac{1}{2}\left(\arcs \xi+\frac{\pi}{2}\right)
392: %\left(\sqrt{\gamo^2-1}+\sqrt{\gamt^2-1}\right) +
393: %\frac{\pi}{2}(1-\gamo) \bigg], \quad \frac{\theta}{\theta+1}\le
394: %\alpha \le \frac{1}{2}, \\
395: %h_{\theta,\alpha}(u) &=& (1+\theta)\sqrt{2}/2 -
396: %h_{\theta,1-\alpha}((1-\theta)\sqrt{2}/2 - u),
397: %\qquad\qquad\qquad\qquad\qquad \frac{1}{2} < \alpha < 1, \\
398: %\xi &=&
399: %\frac{u-\alpha(1-\theta)\sqrt{2}/2}{\sqrt{2\theta\alpha(1-\alpha)}},
400: %\qquad u \in [-\beta_1(\theta,\alpha),\beta_2(\theta,\alpha)], \\
401: %\gamo &=&
402: %\frac{\alpha+\theta(1-\alpha)}{2\sqrt{\theta\alpha(1-\alpha)}}, \\
403: %\qquad
404: %\gamt &=&
405: %\frac{\theta\alpha+1-\alpha}{2\sqrt{\theta\alpha(1-\alpha)}}, \\
406: %\end{eqnarray*}
407: %
408: %
409: %
410: \begin{eqnarray*}
411: \overline{\beta} &=& \sqrt{2\theta\alpha(1-\alpha)}, \\
412: \beta_1 &=& \overline{\beta} - \alpha(1-\theta)\sqrt{2}/2, \qquad
413: \beta_2\ \ =\ \ \overline{\beta} + \alpha(1-\theta)\sqrt{2}/2, \\
414: h_{\theta,\alpha}(u) &=& \theta\sqrt{2}/2 \pm
415: (\beta_1 - \theta\sqrt{2}/2) +
416: \frac{2\overline{\beta}}{\pi} \bigg[ \pm (-\xi-\gamma_1) \arct
417: \sqrt{\frac{(1-\xi)(\gamma_1-1)}{(1+\xi)(\gamma_1+1)}} \\
418: & & +(\xi-\gamma_2)\arct
419: \sqrt{\frac{(1+\xi)(\gamma_2-1)}{(1-\xi)(\gamma_2+1)}} \\ & &
420: + \frac{1}{2}\left( \arcs \xi+\frac{\pi}{2}\right)
421: \frac{1-\theta}{\sqrt{2}\, \overline{\beta}} \pm
422: \frac{\pi}{2}(\gamma_1-1) \bigg], \quad\qquad 0<\alpha\le \frac{1}{2},\\
423: \pm &=& \left\{
424: \begin{array}{ll} + & \quad 0 < \alpha \le \theta/(1+\theta), \\
425: - & \quad \theta/(1+\theta) < \alpha \le 1/2,
426: \end{array} \right. \\
427: \xi &=& \frac{u-\alpha(1-\theta)\sqrt{2}/2}{\overline{\beta}}, \qquad
428: u \in [-\beta_1,\beta_2], \\
429: \gamma_1 &=& \frac{\alpha+\theta(1-\alpha)}{\sqrt{2}\,\overline{\beta}},
430: \qquad
431: \gamma_2\ \ =\ \ \frac{\theta\alpha+1-\alpha}{\sqrt{2}\,\overline{\beta}},
432: \\
433: h_{\theta,\alpha}(u) &=& (1+\theta)\sqrt{2}/2 -
434: h_{\theta,1-\alpha}((1-\theta)\sqrt{2}/2-u), \qquad \frac{1}{2} <
435: \alpha <1,
436: \end{eqnarray*}
437: see Figure 3. Set $$M_\theta(x,y) = -\log(L_\theta(x,y)).$$
438:
439: \paragraph{Theorem 5.} For integers $n,m>0$, let ${\cal T}_{n,m}$ be
440: the set of tableaux whose shape is an $n\times m$ rectangular diagram,
441: and let $\pr_{n,m}$ be the uniform probability measure on
442: $\pr_{n,m}$. If
443: $T=(t_{i,j})_{i,j}\in{\cal T}_{n,m}$, define the rescaled tableau
444: surface of $T$ as the function
445: $\tilde{S}_T:[0,1)\times[0,m/n)\to[0,1]$ given by
446: $$ \tilde{S}_T(x,y) = \frac{1}{n m}t_{\lfloor n x\rfloor+1,
447: \lfloor n y \rfloor+1}. $$
448: Let $0<\theta\le 1$. If $m_n$ is a sequence of integers such
449: that $m_n/n \to\theta$ as $n\to\infty$, then for all $\epsilon>0,\
450: x\in[0,1), y\in [0,\theta)$,
451: $$ \pr_{n,m_n}(T\in{\cal T}_{n,m_n} : |\tilde{S}_T(x,y)-L_\theta(x,y)|
452: > \epsilon ) \xrightarrow[n\to\infty]{} 0. $$
453:
454: \begin{figure}
455: \begin{center}
456: \includegraphics{rectfig.eps}
457: \caption{The curves $h_{\theta,\alpha}$ for $\theta=0.5$,
458: $\alpha=k/9$, $k=1,2,\ldots,8$.}
459: \end{center}
460: \end{figure}
461:
462: \paragraph{Theorem 6.} For each $\theta>0$, $M_\theta$ is the limit
463: surface of uniform random plane partitions of $m_n$ over a rectangular
464: diagram of sides $n$ and $k_n$, provided $k_n/n \to \theta$ and
465: $m_n/n^6\to\infty$ as $n\to\infty$. (The precise statement is by
466: analogy with Theorems 1, 4, and 5.)
467:
468: \subsection{Organization of the paper}
469:
470: The remainder of the paper is organized as follows: In the next
471: section, we present the variational approach to the limit surface of
472: random square Young tableaux, based on the hook formula of
473: Frame-Thrall-Robinson. The level curves of $L$ appear as minimizers of
474: a certain functional. This leads to a proof of Theorem 1 in the
475: interior of the square, except for the explicit identification of
476: $L$. Section 3 is dedicated to the derivation of the explicit formula
477: for the minimizer. Unlike the analogous results of Logan-Shepp,
478: Vershik-Kerov and Cohn-Larsen-Propp, we will show that there is no
479: need to guess the minimizer, by giving a technique for systematically
480: deriving it using an inversion formula for Hilbert transforms on a
481: finite interval. This may prove useful in similar problems.
482:
483: In section 4, we complete the proof of Theorem 1, treating the more
484: delicate case of the boundary of the square, and prove Theorem 3. In
485: section 5, we discuss the \emph{hook walk} of Greene-Nijenhuis-Wilf
486: and the concept of the \emph{co-transition measure} of a Young
487: diagram. Using the explicit formulas for the co-transition measure
488: derived in \cite{romik}, we compute the co-transition measure of
489: the level curves $g_\alpha$, proving Theorem 2. In section 6, we prove
490: Theorem 4. In section 7 we give the computations necessary for
491: settling the rectangular case. In section 8, we discuss the
492: connections of our results to the theory of Plancherel-random
493: partitions, and some open problems.
494:
495: \section{A variational problem for random square tableaux}
496:
497: \subsection{A large-deviation principle}
498: One may consider a tableau $T\in{\cal T}_n$ as a path in the
499: \emph{Young graph} of all Young diagrams, starting with the empty
500: diagram, and leading up to the $n\times n$ square diagram, where each
501: step is of adding one box to the diagram. Identify $T$ with this
502: sequence $\lambda_T^0=\phi \subset \lambda_T^1 \subset \lambda_T^2
503: \subset \ldots \subset \lambda_T^{n^2} = \square_n$ of
504: diagrams. ($\lambda_T^k$ is simply the sub-diagram of the square
505: comprised of those boxes where the value of the entry of $T$ is $\le
506: k$.) Theorem 1 is then roughly equivalent, in a sense that will be
507: made precise later, to the statement that for each $1\le k\le n^2-1$, the
508: rescaled shape of $\lambda_T^k$ for a random
509: $T\in{\cal T}_n$ resembles the \emph{sub-level set}
510: $$ \{(x,y)\in[0,1]^2 : L(x,y) \le k/n^2 \} $$ of $L$, with
511: probability $1-o(1)$ as $n\to\infty$. It is this approach that leads
512: to the large-deviation principle. Namely, we can estimate the
513: probability that the sub-diagram $\lambda_T^k$ has a given shape:
514:
515: \paragraph{Lemma 1.} For $T\in{\cal T}_n$, denote as before
516: $\lambda_T^0 \subset \ldots \subset \lambda_T^{n^2}$ the path in the
517: Young graph defined by $T$, and for each $0\le k\le n^2$, let
518: $\lambda_T^k : \lambda_T^k(1) \ge \lambda_T^k(2) \ge \ldots\ge
519: \lambda_T^k(n)$ be the lengths of the columns of $\lambda_T^k$ (some
520: of them may be $0$). For any Young diagram $\lambda:\lambda(1)\ge
521: \lambda(2)\ge \ldots\ge \lambda(n)$ whose graph lies within the $n\times n$
522: square, define the function $f_{\lambda}:[0,1]\to[0,1]$ by
523: \begin{equation}\label{eq:lambdagraph}
524: f_{\lambda}(x) = \frac{1}{n}\lambda(\lceil n x\rceil).
525: \end{equation}
526: (Note that this depends implicitly on $n$.) Let $0\le k \le n^2$, and
527: let $\alpha=k/n^2$. Then for any given diagram $\lambda_0 \subseteq
528: \square_n$ with area $k$, we have
529: \begin{equation}\label{eq:lemma1}
530: \pr_n\left( T\in{\cal T}_n : \lambda_T^k =
531: \lambda_0 \right) =
532: \exp\bigg(-(1+o(1)) n^2 (I(f_{\lambda_0})+H(\alpha)+C)\bigg)
533: \end{equation}
534: as $n\to\infty$, where
535: \begin{eqnarray*}
536: C &=& \frac{3}{2}-2\log 2, \\
537: H(\alpha) &=& -\alpha\log(\alpha)-(1-\alpha)\log(1-\alpha), \\
538: I(g) &=& \int_0^1 \int_0^1 \log|g(x)-y+g^{-1}(y)-x|dy\, dx, \\
539: g^{-1}(y) &=& \inf\{x\in[0,1]: g(x)\le y\}.
540: \end{eqnarray*}
541: The $o(1)$ is uniform over all $\lambda_0$ and all $0\le k\le n^2$.
542:
543: \paragraph{Proof.} For a Young diagram $\lambda:\lambda(1)\ge
544: \lambda(2)\ge \ldots \ge \lambda(l)$ of area $m$, denote by $d(\lambda)$
545: the number of Young tableaux of shape $\lambda$ (also known as the
546: \emph{dimension} of $\lambda$, as it is known to be equal to the
547: dimension of a certain irreducible representation corresponding to
548: $\lambda$ of the symmetric group of order $m$). Recall the \emph{hook
549: formula} of Frame-Thrall-Robinson \cite{framehook}, which says that
550: $d(\lambda)$ is given by
551: \begin{equation}\label{eq:hook}
552: d(\lambda) = \frac{m!}{\prod_{(i,j)\in \lambda} h_{i,j}},
553: \end{equation}
554: where the product is over all boxes $(i,j)$ in the diagram, and
555: $h_{i,j}$ is the \emph{hook number} of a box, given by
556: \begin{eqnarray*}
557: h_{i,j} &=& \lambda(i) - j + \lambda'(j) - i + 1\\ &=& 1+\textrm{number of
558: boxes either to the right of, or below $(i,j)$}
559: \end{eqnarray*}
560: (and where $\lambda'$ is the conjugate partition to $\lambda$.)
561: Then we have
562: \footnote{Note to the reader: this is probably the most important
563: formula in the paper!}
564: \begin{equation}\label{eq:dmeasure}
565: \pr_n\left( T\in{\cal T}_n : \lambda_T^k =
566: \lambda_0 \right) = \frac{d(\lambda_0)
567: d(\square_n\setminus \lambda_0)}{d(\square_n)},
568: \end{equation}
569: where $d(\square_n\setminus \lambda_0)$ means the number of fillings
570: of the numbers $1,\ldots,n^2-k$ in the cells of the skew-Young diagram
571: $\square_n\setminus \lambda_0$ that are monotonically
572: \emph{decreasing} along rows and columns. This is because
573: $\square_n\setminus\lambda_0$ can be thought of as an ordinary
574: diagram, when viewed from the opposite corner of the square. The
575: number of square tableaux whose $k$-th subtableau has shape
576: $\lambda_0$ is simply the number of tableaux of shape $\lambda_0$,
577: times the number of fillings of the numbers $k+1,k+2,\ldots,n^2$ in the
578: cells of $\square_n\setminus\lambda_0$ that are monotonically
579: increasing along rows and columns -- and these are of course
580: isomorphic to tableaux of shape $\square_n\setminus\lambda_0$, by
581: replacing each entry $i$ with $n^2+1-i$.
582:
583: Take minus the logarithm of \eqref{eq:dmeasure} and divide by $n^2$,
584: using \eqref{eq:hook}. The right-hand side becomes
585: $$ a+b+c-d := \frac{1}{n^2}\log \left(\frac{(n^2)!}{k!(n^2-k)!}\right)
586: + \frac{1}{n^2}\sum_{i=1}^n \sum_{j=1}^{\lambda(i)} \log(\lambda(i) -
587: j + \lambda'(j)-i+1) $$ $$ + \frac{1}{n^2}\sum_{i=1}^n
588: \sum_{j=\lambda(i)+1}^n \log(j-\lambda(i)+i-\lambda'(j)+1) -
589: \frac{1}{n^2} \sum_{i=1}^n \sum_{j=1}^n \log(2 n - i - j + 1).$$
590: By Stirling's formula, we have $ a = n^{-2} \log\binom{n^2}{k} =
591: H(\alpha)+o(1)$, with the required uniformity in $k$. The other
592: summands look like Riemann sums of double integrals. Indeed, we claim
593: that
594: \begin{eqnarray*}
595: b &=& \int_0^1 \int_0^{f_{\lambda_0}(x)} \log\bigg(
596: f_{\lambda_0}(x)-y+f_{\lambda_0}^{-1}(y)-x \bigg) dy\,dx +
597: \frac{k}{n^2}\log n + o(1), \\
598: c &=& \int_0^1 \int_{f_{\lambda_0}(x)}^1 \log\bigg(
599: y-f_{\lambda_0}(x)+x-f_{\lambda_0}^{-1}(y)\bigg) dy\,dx +
600: \frac{n^2-k}{n^2} \log n + o(1), \\
601: d &=& \int_0^1 \int_0^1 \log(2-x-y)dy\,dx + \log n + o(1)
602: = C + \log n +o(1),
603: \end{eqnarray*}
604: which on summing and exponentiating would give the lemma. Let us
605: prove, for example, the first of these equations. Write
606: \begin{eqnarray*}
607: b &=& \frac{1}{n^2}\sum_{i=1}^n
608: \sum_{j=1}^{\lambda(i)} \log(\lambda(i) - j + \lambda'(j)-i+1) \\ &=&
609: \frac{1}{n^2}\sum_{i=1}^n
610: \sum_{j=1}^{\lambda(i)} \log\left(\frac{\lambda(i) - j +
611: \lambda'(j)-i+1}{n}\right) + \frac{k}{n^2}\log n.
612: \end{eqnarray*}
613: Fix $1\le i\le n$ and $1\le j \le \lambda(i)$. Denote $h =
614: (\lambda(i) - j + \lambda'(j)-i+1)/n$. Approximate $n^{-2} \log h$ in
615: the above sum by the double integral
616: $$ Q := \int_{(i-1)/n}^{i/n} \int_{(j-1)/n}^{j/n} \log \left(
617: f_{\lambda_0}(x)-y + f_{\lambda_0}^{-1}(y)-x \right)dy\,dx.
618: $$
619: A change of variables transforms this (check the definition of
620: $f_{\lambda_0}$) into
621: $$ Q = \int_{-1/2n}^{1/2n} \int_{-1/2n}^{1/2n} \log (x+y+h) dx\, dy.
622: $$
623: %
624: % old version
625: %
626: %By the integral mean value theorem, we have for some $-1<\theta<1$
627: %$$ Q = \frac{1}{n^2}\log\left(h+\frac{\theta}{n}\right) =
628: %\frac{1}{n^2} \log h + \frac{1}{n^2} \log\left(1+\frac{\theta}{n
629: %h}\right). $$
630: %(Note that $h$ may take the values $1/n, 2/n, \ldots, (2n-1)/n$. The
631: %value of $\theta$ depends on $h$, and the integral mean value theorem
632: %guarantees that $-1 \le \theta \le 1$. However, since $\log(x+y+h)$ is
633: %strictly increasing as a function of $x$ and $y$, clearly $\theta \in
634: %(-1,1)$. Special care must be taken when $h=1/n$, where one needs to
635: %check first that $Q > -\infty$.) It follows that
636: %$$ Q = \frac{1}{n^2}\log h + O\left(\frac{1}{n^3}\right). $$
637: %Summing over all $1\le i\le n$, $1\le j\le \lambda(i)$ gives the
638: %equation for $b$.
639: %
640: %
641: % new version 28/03:
642: %
643: Note that $h$ may take the values $1/n, 2/n, \ldots, (2n-1)/n$. If
644: $h=1/n$, then integrating we get
645: $$ Q = -\frac{\log n}{n^2} + n^{-2} \int_0^1 \int_0^1 \log(u+v)du\,dv
646: = \frac{\log h}{n^2} + O(n^{-2}). $$
647: If $h\ge 2/n$, by the integral mean value theorem, we have for some
648: $\eta \in [-1,1]$,
649: $$ Q = \frac{\log(h+\eta n^{-1})}{n^2} = \frac{\log h}{n^2} + O((n^3
650: h)^{-1}). $$
651: Clearly then the last estimate holds for $h=1/n$ as well. The sum of
652: the remainders over all $1 \le i \le n$, $1 \le j \le \lambda(i)$ is
653: of order
654: $$ n^{-2}\sum_{(i,j)\in \lambda_0} \frac{1}{h_{i,j}} \le n^{-2}
655: \sum_{m=1}^{2n-1} \frac{a(m)}{m}, $$
656: where
657: $$ a(m) := \#\{ (i,j)\in \lambda_0 : h_{i,j}=m \}. $$
658: Clearly $a(m) \le n$, since each row $i$ of $\lambda_0$ contains at
659: most one cell $(i,j)$ with $h_{i,j}=m$. This gives that the sum of the
660: remainders is of order
661: $$ n^{-2}\sum_{m=1}^{2n-1} \frac{n}{m} = O\left(\frac{\log
662: n}{n}\right),$$
663: which is indeed $o(1)$.
664: \qed
665:
666: \subsection{Two formulations of the variational problem}
667:
668: Lemma 1 says, roughly, that the exponential order of the probability
669: that a random square tableau $T$ has a given $k$-subtableau shape,
670: where $k$ is approximately $\alpha\cdot n^2$, is given by the value of
671: the functional $I$ on the boundary $g$ of the shape, plus some terms
672: depending only on $\alpha$. Following the well-known methodology of
673: large deviation theory, the natural next step is to identify the
674: global minimum of $I$ over the appropriate class of functions, or in
675: other words to find the \emph{most likely} shape for the
676: $\alpha$-level set. If we can prove that there is a unique minimum,
677: and identify it, that will be a major step towards proving Theorem 1.
678: So we have arrived at the following variational problem.
679:
680: \paragraph{Variational problem 1.} For each $0<\alpha<1$, any weakly
681: decreasing function $f:[0,1]\to[0,1]$ such that $\int_0^1
682: f(x)dx=\alpha$ is called \emph{$\alpha$-admissible}. Find the unique
683: $\alpha$-admissible function that minimizes the functional
684: $$ I(f) = \int_0^1 \int_0^1 \log|f(x)-y+f^{-1}(y)-x|dy\, dx. $$
685:
686: \bigskip
687: We now simplify the form of the functional $I$, by first rotating the
688: coordinate axes by 45 degrees, and then reparametrizing the square by
689: the ``hook coordinates'' -- an idea used in \cite{vershikkerov1},
690: \cite{vershikkerov2}, \cite{loganshepp}. Let $u, v$ be the rotated
691: coordinates as in \eqref{eq:rotated}. Given an $\alpha$-admissible
692: function $f:[0,1]\to[0,1]$, there corresponds to it a function
693: $g:[-\sqrt{2}/2, \sqrt{2}/2]\to[0,\sqrt{2}]$, such that
694: $$ y = f(x) \iff v = g(u) $$
695: (see Figure 4). The class of $\alpha$-admissible functions translates
696: to those functions $g:[-\sqrt{2}/2,\sqrt{2}/2]\to[0,\sqrt{2}]$ that
697: are $1$-Lipschitz, and satisfy
698: $g(-\sqrt{2}/2)=g(\sqrt{2}/2)=\sqrt{2}/2 $ and
699: \begin{equation}\label{eq:alphacond}
700: \int_{-\sqrt{2}/2}^{\sqrt{2}/2} (g(u)-|u|)du = \alpha.
701: \end{equation}
702: We continue to call such functions $\alpha$-admissible. We call a
703: function admissible if it is $\alpha$-admissible for some $0\le \alpha
704: \le 1$.
705:
706: \begin{figure}[h!]
707: \begin{center}
708: \begin{picture}(500,160)(0,0)
709: \put(20,20){\vector(1,0){120}}
710: \put(20,20){\vector(0,1){120}}
711: \put(20,20){\framebox(100,100)}
712: \put(10,135){$y$}
713: \put(14,120){\line(1,0){20}}
714: \put(8,116){\footnotesize 1}
715: \put(135,10){$x$}
716: \put(120,14){\line(0,1){20}}
717: \put(118,5){\footnotesize 1}
718: \put(140,70){\vector(1,0){30}}
719: \put(170,70){\vector(-1,0){30}}
720: \put(180,20){\vector(1,0){190}}
721: \put(270,20){\line(-1,1){70}}
722: \put(270,20){\line(1,1){70}}
723: \put(200,90){\line(1,1){70}}
724: \put(340,90){\line(-1,1){70}}
725: \multiput(340,20)(0,5){14}{\circle*{1}}
726: \multiput(200,20)(0,5){14}{\circle*{1}}
727: \put(180,20){\vector(0,1){120}}
728: \put(360,10){$u$}
729: \put(170,135){$v$}
730: \put(270,20){\line(0,-1){5}}
731: \put(269,3){0}
732: \put(185,3){\footnotesize $-\sqrt{2}/2$}
733: \put(328,3){\footnotesize $\sqrt{2}/2$}
734: \put(200,20){\line(0,-1){5}}
735: \put(340,20){\line(0,-1){5}}
736: %
737: \qbezier(20,110)(30,85)(50,60)
738: \qbezier(50,60)(85,25)(120,20)
739: \qbezier(207,83)(232,72)(254,69)
740: \qbezier(254,69)(296,62)(340,90)
741: \put(280,92){\circle*{3}}
742: \put(270,97){\footnotesize $(u,v)$}
743: \put(78,64){\circle*{3}}
744: \put(70,70){\footnotesize $(x,y)$}
745: \put(280,92){\line(-1,-1){23}}
746: \put(280,92){\line(1,-1){20}}
747: \put(78,64){\line(-1,0){30}}
748: \put(78,64){\line(0,-1){27}}
749: \multiput(257,69)(0,-3){17}{\circle*{1}}
750: \multiput(300,72)(0,-3){18}{\circle*{1}}
751: \put(300,20){\line(0,-1){5}}
752: \put(257,20){\line(0,-1){5}}
753: \put(299,3){\footnotesize $t$}
754: \put(256,3){\footnotesize $s$}
755: \put(33,95){\footnotesize $f(x)$}
756: \put(221,84){\footnotesize $g(u)$}
757: \end{picture}
758: \caption{The rotated graph and the hook coordinates $s,t$}
759: \end{center}
760: \end{figure}
761:
762: To derive the new form of the functional, write
763: $$ I(f) = I_1(f) + I_2(f) := \int_0^1 \int_0^{f(x)}
764: \log(h_f(x,y))dy\,dx + \int_0^1 \int_{f(x)}^1 \log(h_f(x,y))dy\,dx, $$
765: where $h_f(x,y)$ is the \emph{hook function} of $f$,
766: $$ h_f(x,y) = |f(x)-y + f^{-1}(y)-x|. $$
767: Now, set $$J(g) = J_1(g) + J_2(g) := I_1(f) + I_2(f),$$
768: where $f$ and $g$ are rotated versions of the same graph as in Figure
769: 4. Then
770: $$ J_2(g) = \int_{-\sqrt{2}/2}^{\sqrt{2}/2} \int_{g(u)}^{\sqrt{2}-|u|}
771: \log h_f(x,y) dv\,du. $$
772: Reparametrize this double integral by the \emph{hook coordinates} $s$
773: and $t$,
774: $$ s = \frac{f^{-1}(y)-y}{\sqrt{2}}, \qquad t =
775: \frac{x-f(x)}{\sqrt{2}} $$
776: (see Figure 4). The Lipschitz property ensures that this
777: transformation is one-to-one from the region
778: $$ \{ (u,v) : -\sqrt{2}/2 \le u \le \sqrt{2}/2,\ g(u)\le v \le
779: \sqrt{2}-|u| \} $$
780: onto the region
781: $$ \Delta = \{(s,t) : -\sqrt{2}/2 \le s \le t \le \sqrt{2}/2 \}. $$
782: Therefore the integral transforms as
783: $$ J_2(f) = \int\!\!\int_\Delta \log\left(\sqrt{2}(t-s)\right)
784: \left|\frac{\partial(u,v)}{\partial(s,t)}\right|ds\,dt.
785: $$
786: It remains to compute the Jacobian $\partial(u,v)/\partial(s,t)$. An
787: easy computation gives (see \cite{vershikkerov1},
788: \cite{vershikkerov2}, \cite{loganshepp})
789: $$ \frac{\partial(u,v)}{\partial(s,t)} =
790: \frac{1}{2}(1-g'(s))(1+g'(t)). $$
791: (This can be viewed geometrically as follows: draw on the $u$-axis in
792: Figure 4 the two intervals $[s,s+ds], [t,t+dt]$. The set of points in
793: the square for which the hook coordinates fall inside the two
794: intervals is approximately a parallelogram whose area is clearly seen
795: from the picture to be linear in $1-g'(s)$ and in $1+g'(t)$.)
796: So
797: $$ J_2(g) = \frac{1}{2}\int\!\!\int_\Delta
798: \log\left(\sqrt{2}(t-s)\right) (1-g'(s))(1+g'(t)) ds\,dt. $$
799: A similar computation for $J_1$, using ``lower'' instead of ``upper''
800: hook coordinates, shows that
801: $$ J_1(g) = \frac{1}{2}\int\!\!\int_\Delta
802: \log\left(\sqrt{2}(t-s)\right) (1+g'(s))(1-g'(t)) ds\,dt. $$
803: This gives
804: \begin{eqnarray*}
805: J(g)&=& \frac{1}{2}\int\!\!\int_\Delta \log\left(\sqrt{2}(t-s)\right)
806: \big[(1-g'(s))(1+g'(t))+(1+g'(s))(1-g'(t))\big]ds\,dt \\
807: &=& \frac{1}{2}
808: \int\!\!\int_\Delta
809: \log\left(\sqrt{2}(t-s)\right)\big(2-2g'(s)g'(t)\big)ds\,dt \\
810: &=& -\frac{1}{2}
811: \int_{-\sqrt{2}/2}^{\sqrt{2}/2}\int_{-\sqrt{2}/2}^{\sqrt{2}/2}
812: \log|t-s|\cdot g'(s)g'(t)ds\,dt + \log 2-\frac{3}{2}.
813: \end{eqnarray*}
814: We can now state a reformulation of the original variational problem.
815:
816: \paragraph{Variational problem 2.} For each $0<\alpha<1$, a function
817: $g:[-\sqrt{2}/2,\sqrt{2}/2]\to[0,\sqrt{2}]$ is called
818: $\alpha$-admissible if: $g(-\sqrt{2}/2)=g(\sqrt{2}/2)=\sqrt{2}/2$; $g$
819: is 1-Lipschitz; and $\int_{-\sqrt{2}/2}^{\sqrt{2}/2}
820: (g(u)-|u|)du = \alpha$. Find the unique $\alpha$-admissible function
821: that minimizes the functional
822: \begin{equation}\label{eq:functionalk}
823: K(g) = -\frac{1}{2}\int_{\sqrt{2}/2}^{\sqrt{2}/2}
824: \int_{-\sqrt{2}/2}^{\sqrt{2}/2} g'(s)g'(t) \log|s-t| ds dt.
825: \end{equation}
826:
827: \subsection{Deduction of Theorem 1(ii)}
828:
829: In the next section, we prove the following theorem.
830:
831: \paragraph{Theorem 7.} For each $0<\alpha<1$, let $\tilde{g}_\alpha$
832: be the unique extension of $g_\alpha$ (defined in
833: \eqref{eq:minimizers}) to an $\alpha$-admissible function, namely
834: $$ \tilde{g}_\alpha(u) = \left\{ \begin{array}{ll}
835: g_\alpha(u) & |u|\le \sqrt{2\alpha(1-\alpha)} \\
836: |u| & \sqrt{2\alpha(1-\alpha)} \le |u| \le \sqrt{2}/2 \end{array}
837: \right. $$
838: for $0<\alpha\le 1/2$, and
839: $$ \tilde{g}_\alpha(u) = \left\{ \begin{array}{ll} g_\alpha(u) &
840: |u|\le \sqrt{2\alpha(1-\alpha)} \\ \sqrt{2}-|u| &
841: \sqrt{2\alpha(1-\alpha)} \le |u| \le \sqrt{2}/2 \end{array} \right. $$
842: for $1/2 < \alpha < 1$.
843: Then:
844:
845: (i) $\tilde{g}_\alpha$ is the unique solution to Variational
846: problem 2;
847:
848: (ii) $K(\tilde{g}_\alpha) = -H(\alpha)+\log 2$;
849:
850: (iii) For any $\alpha$-admissible function $g$ we have
851: $$ K(g) \ge K(\tilde{g}_\alpha) + K(g-\tilde{g}_\alpha). $$
852:
853: \bigskip
854: Assuming this as proven, our goal is now to prove Theorem 1. At the
855: beginning of this section, we claimed that Theorem 1 was equivalent to
856: the statement that the subtableau $\lambda_T^k$ has shape
857: approximately described by the region bounded under the graph of the
858: level curve $\{L=k/n^2\}$ (which in rotated coordinates is given by
859: the curve $v=\tilde{g}_\alpha(u)$, where $\alpha=k/n^2$). We shall now
860: make precise the sense in which this is true, and see how this follows
861: from the fact that $\tilde{g}_\alpha$ is the minimizer.
862:
863: For a continuous function $p:[-\sqrt{2}/2,\sqrt{2}/2]\to\mathbb{R}$,
864: define its supremum norm
865: $$ ||p||_\infty = \max_{u\in[-\sqrt{2}/2,\sqrt{2}/2]}
866: |p(u)|. $$
867:
868: \paragraph{Lemma 2.} $K$ is continuous in the supremum norm on the
869: space of admissible functions.
870:
871: \paragraph{Proof.} Consider the symmetric bilinear form
872: \begin{equation}\label{eq:bilinear}
873: \langle g,h \rangle = -\frac{1}{2}\int_{-\sqrt{2}/2}^{\sqrt{2}/2}
874: \int_{-\sqrt{2}/2}^{\sqrt{2}/2} g'(s)h'(t) \log|s-t|ds\,dt
875: \end{equation}
876: defined whenever $g$ and $h$ are almost everywhere differentiable
877: functions on $[-\sqrt{2}/2,\sqrt{2}/2]$ with bounded derivative. We
878: show that $\langle \cdot,\cdot\rangle$ is continuous in the supremum
879: norm with respect to any of its arguments, when restricted to the set
880: of $1$-Lipschitz functions; this will imply the lemma, since $K(g) =
881: \langle g,g\rangle$.
882: %
883: %Integrating \eqref{eq:bilinear} by
884: %parts gives
885: %\begin{equation}\label{eq:scalarprod}
886: %\langle g,h \rangle = -\frac{\pi}{2}\int_{-\sqrt{2}/2}^{\sqrt{2}/2}
887: %g'(s)\tilde{h}(s)ds,
888: %\end{equation}
889: %where $\tilde{h}$ is the Hilbert transform of $h$,
890: %$$ \tilde{h}(s) = \frac{1}{\pi} \int_{-\sqrt{2}/2}^{\sqrt{2}/2}
891: %\frac{h(t)}{t-s}dt$$
892: %(This is easy to justify formally given the restrictions on $g,h$.)
893: %
894: Write \eqref{eq:bilinear} more carefully as
895: \begin{equation*}
896: \langle g,h \rangle = -\frac{1}{2}\int_{-\sqrt{2}/2}^{\sqrt{2}/2}
897: g'(s) \cdot \lim_{\epsilon\searrow 0} \left[
898: \int_{-\sqrt{2}/2}^{s-\epsilon} h'(t) \log(s-t)dt +
899: \int_{s+\epsilon}^{\sqrt{2}/2} h'(t) \log(t-s)dt \right] ds.
900: \end{equation*}
901: For $s\in (-\sqrt{2}/2,\sqrt{2}/2)$ which is a point of
902: differentiability of $h$, integration by parts gives
903: \begin{multline*}
904: \int_{-\sqrt{2}/2}^{s-\epsilon} h'(t) \log(s-t)dt +
905: \int_{s+\epsilon}^{\sqrt{2}/2} h'(t) \log(t-s)dt = \\ =
906: h(t)\log(s-t)\bigg|_{t=-\sqrt{2}/2}^{t=s-\epsilon} -
907: \int_{-\sqrt{2}/2}^{s-\epsilon}\frac{h(t)}{t-s}dt +
908: h(t)\log(t-s)\bigg|_{t=s+\epsilon}^{t=\sqrt{2}/2} -
909: \int_{s+\epsilon}^{\sqrt{2}/2} \frac{h(t)}{t-s}dt \\
910: = h\left(\frac{\sqrt{2}}{2}\right)\log\left(\frac{\sqrt{2}}{2}-s\right)
911: -h\left(-\frac{\sqrt{2}}{2}\right)\log\left(\frac{\sqrt{2}}{2}+s\right)
912: \qquad\qquad\qquad\qquad\qquad\qquad\ \
913: \\ +(h(s-\epsilon)-h(s+\epsilon))\log\epsilon -
914: \int_{[-\sqrt{2}/2,s-\epsilon]\cup [s+\epsilon,\sqrt{2}/2]}
915: \frac{h(t)}{t-s}dt
916: \qquad\qquad\qquad\ \ \
917: \\ \xrightarrow[\ \ \ \epsilon \searrow 0\ \ \ ]{}
918: h\left(\frac{\sqrt{2}}{2}\right)\log\left(\frac{\sqrt{2}}{2}-s\right)
919: -h\left(-\frac{\sqrt{2}}{2}\right)\log\left(\frac{\sqrt{2}}{2}+s\right)
920: - \pi\tilde{h}(s),
921: \qquad\quad\ \
922: \end{multline*}
923: where $\tilde{h}$ is the Hilbert transform of $h$, defined by the
924: principal value integral
925: $$ \tilde{h}(s) = \frac{1}{\pi}\int_{\mathbb{R}} \frac{h(t)}{t-s}dt $$
926: (think of $h$ as a function on $\mathbb{R}$ which is $0$ outside
927: $[-\sqrt{2}/2,\sqrt{2}/2]$.) Going back to \eqref{eq:bilinear}, this gives
928:
929: \begin{multline}\label{eq:scalarprod}
930: \qquad \langle g,h\rangle = -\frac{1}{2}h\left(\frac{\sqrt{2}}{2}\right)
931: \int_{-\sqrt{2}/2}^{\sqrt{2}/2}
932: g'(s)\log\left(\frac{\sqrt{2}}{2}-s\right) ds \\ +
933: \frac{1}{2}h\left(\frac{-\sqrt{2}}{2}\right)
934: \int_{-\sqrt{2}/2}^{\sqrt{2}/2}
935: g'(s)\log\left(\frac{\sqrt{2}}{2}+s\right) ds
936: \qquad\qquad\qquad \ \\ +
937: \frac{\pi}{2} \int_{-\sqrt{2}/2}^{\sqrt{2}/2} g'(s)\tilde{h}(s)ds.
938: \qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\quad\
939: \end{multline}
940: Now recalling that the Hilbert transform is an isometry on
941: $L_2(\mathbb{R})$ (see \cite{titchmarsh}, Theorem 90), and using the
942: fact that
943: $$ \bigg| \int_{-\sqrt{2}/2}^{\sqrt{2}/2} \log\left( \frac{\sqrt{2}}{2}\pm
944: s\right) ds \bigg| = \frac{2-\log 2}{\sqrt{2}} < 1,$$
945: this implies that for 1-Lipschitz functions $g,h_1,h_2$,
946: \begin{eqnarray*}
947: |\langle g,h_1-h_2\rangle| &\le& ||h_1-h_2||_\infty +
948: \frac{\pi}{2} \int_{-\sqrt{2}/2}^{\sqrt{2}/2}
949: |\tilde{h}_1(s)-\tilde{h}_2(s)|ds \\ &\le& ||h_1-h_2||_\infty +
950: 2^{1/4}\frac{\pi}{2} \left(
951: \int_{-\sqrt{2}/2}^{\sqrt{2}/2}
952: \left(\tilde{h}_1(s)-\tilde{h}_2(s)\right)^2ds
953: \right)^{1/2} \\ &\le& ||h_1-h_2||_\infty +
954: 2^{1/4}\frac{\pi}{2} \left( \int_{\mathbb{R}}
955: \left(\tilde{h}_1(s)-\tilde{h}_2(s)\right)^2ds \right)^{1/2} \\ &=&
956: ||h_1-h_2||_\infty + 2^{1/4}\frac{\pi}{2}
957: \left(\int_{-\sqrt{2}/2}^{\sqrt{2}/2}(h_1(s)-h_2(s))^2ds
958: \right)^{1/2} \\ &\le&
959: \left(1+2^{1/2} \frac{\pi}{2}\right) ||h_1-h_2||_\infty.
960: \end{eqnarray*}
961:
962: \vspace{-26.0pt}
963: \qed
964:
965: \bigskip\medskip
966: We have another use for \eqref{eq:scalarprod}. Let $f$ be a Lipschitz
967: function on $[-\sqrt{2}/2,\sqrt{2}/2]$ that satisfies $f(\pm
968: \sqrt{2}/2)=0$. Denote by
969: $$ F[f](x) = \int_{\mathbb{R}} f(t)e^{-ixt}dt $$
970: the Fourier transform of a function $f$. Recall the well-known
971: formulas
972: \begin{eqnarray*}
973: F[\tilde{f}](x) &=& i\cdot \sgn x\cdot F[f](x),\\
974: F[f'](x) &=& i\cdot x \cdot F[f](x),\\
975: \int_{\mathbb{R}} f_1(t)\overline{f_2(t)}dt &=& \frac{1}{2\pi}
976: \int_{\mathbb{R}}F[f_1](x) \overline{F[f_2](x)} dx.
977: \end{eqnarray*}
978: Then, by \eqref{eq:scalarprod}
979: \begin{multline}\label{eq:fourierk}
980: \qquad\qquad\ \,\,\,
981: K(f) = \langle f,f\rangle =
982: \frac{\pi}{2} \int_{-\sqrt{2}/2}^{\sqrt{2}/2} f'(s)\tilde{f}(s)ds
983: \\ =
984: \frac{1}{4}\int_{\mathbb{R}} F[f'](x)\overline{F[\tilde{f}](x)}dx =
985: \frac{1}{4}\int_{\mathbb{R}} |x|\cdot|F[f](x)|^2dx.
986: \qquad\qquad\qquad
987: \end{multline}
988: We note as a lemma an important consequence of this identity which we
989: shall need later on.
990:
991: \paragraph{Lemma 3.} If $f$ is a Lipschitz function with $f(\pm
992: \sqrt{2}/2)=0$ as above, then $K(f)\ge 0$, and $K(f)=0$ only if
993: $f\equiv 0$. \qed
994:
995: \bigskip
996: Lemma 3 will be used in the next section to easily deduce uniqueness
997: of the minimizer. In fact, Theorem 7 gives all the necessary
998: information to prove a non-quantitative version of Theorem 1,
999: i.e. without the rate-of-convergence estimates. However, we can do
1000: better, by noting that Theorem 7(iii), together with the
1001: representation \eqref{eq:fourierk}, can be used to give
1002: quantitative estimates for the rate of convergence in Theorem 1. We
1003: prove the following strengthening of Lemma 3:
1004:
1005: \paragraph{Lemma 4.} For every $r\in(2,3)$,
1006: there exists a constant $c=c(r) > 0$ such that for all 2-Lipschitz
1007: functions $f:[-\sqrt{2}/2,\sqrt{2}/2]\to \mathbb{R}$ that satisfy
1008: $f(\pm \sqrt{2}/2) = 0$, we have
1009: $$ K(f) \ge c ||f||_\infty^r. $$
1010:
1011: \paragraph{Proof.} Had the power of $|x|$ in \eqref{eq:fourierk} been
1012: 2, $K(f)$ would have been equal to $1/4$ times the squared $L_2$-norm
1013: of $x F[f](x) = F[f'](x)$. Having $|x|$ in
1014: \eqref{eq:fourierk} invites the conclusion that instead we are dealing
1015: with the squared $L_2$-norm of $f^{(1/2)}(x)$, the \emph{fractional}
1016: derivative of $f$ of order $1/2$.
1017:
1018: To see that this is indeed the case, and to use the full power of such
1019: an interpretation of $K(f)$, let us recall the corresponding
1020: definitions. For $\alpha \in (0,1)$, the fractional derivative
1021: $f^{(\alpha)}(x)$ of order $\alpha$ is defined by
1022: \begin{equation}\label{eq:fractional}
1023: f^{(\alpha)}(x) = \frac{\alpha}{\Gamma(1-\alpha)} \int_0^\infty
1024: \frac{f(x)-f(x-t)}{t^{1+\alpha}} dt.
1025: \end{equation}
1026: The integral exists as $f(x)$ is Lipschitz and bounded. Clearly
1027: $f^{(\alpha)}(x)\equiv 0$ for $x\le -\sqrt{2}/2$. Then
1028: \begin{multline}\label{eq:b4}
1029: \qquad\qquad
1030: F[f^{(\alpha)}](x) = \int_\mathbb{R} e^{-ixt}f^{(\alpha)}(t)dt
1031: \\ = \frac{\alpha}{\Gamma(1-\alpha)} \int_0^\infty
1032: \frac{1-e^{-ix\tau}}{\tau^{1+\alpha}} d\tau \cdot F[f](x)
1033: = (ix)^{\alpha} F[f](x), \qquad\ \
1034: \end{multline}
1035: where
1036: $$
1037: (ix)^\alpha := \left\{ \begin{array}{ll}
1038: |x|^\alpha \exp(i\alpha\pi/2), & x>0, \\
1039: |x|^\alpha \exp(-i\alpha\pi/2),& x<0. \end{array}\right.
1040: $$
1041: Indeed, setting
1042: $$ z^{1+\alpha} = |z|\exp(i(1+\alpha)\theta)),
1043: \quad \textrm{if }z=|z|e^{i\theta},\ \ \theta\in(-\pi,\pi), $$
1044: we have
1045: \begin{eqnarray*}
1046: \int_0^\infty \frac{1-e^{-ix\tau}}{\tau^{1+\alpha}}d\tau &=&
1047: (ix)^\alpha \int_0^{i\infty} \frac{1-e^{-z}}{z^{1+\alpha}}dz =
1048: (ix)^\alpha \int_0^\infty \frac{1-e^{-\tau}}{\tau^{1+\alpha}}d\tau \\
1049: &=& (ix)^\alpha \frac{1}{\alpha} \int_0^\infty
1050: \tau^{-\alpha}e^{-\tau}d\tau = (ix)^\alpha
1051: \frac{\Gamma(1-\alpha)}{\alpha}.
1052: \end{eqnarray*}
1053: In particular, for $\alpha=1/2$, we get from \eqref{eq:b4} that
1054: $$ \big|F[f^{(1/2)}](x)\big|^2 = |x|\cdot |F[f](x)|^2, $$
1055: whence, by \eqref{eq:b4} and isometry of the Fourier transform,
1056: \begin{equation} \label{eq:b5}
1057: K(f) = \frac{1}{4}\int_\mathbb{R} |x|\cdot |F[f](x)|^2dx =
1058: \frac{\pi}{2}|f^{(1/2)}(x)|^2.
1059: \end{equation}
1060: The fractional integration operator, inverse to that in
1061: \eqref{eq:fractional}, is known to be given by
1062: \begin{equation}\label{eq:b6}
1063: f(x) = (I_\alpha f^{(\alpha)})(x), \qquad
1064: (I_\alpha h)(x) := \frac{1}{\Gamma(\alpha)}\int_{-\infty}^x
1065: (x-t)^{\alpha-1} h(t)dt.
1066: \end{equation}
1067: As a check, the Fourier transform of the RHS is
1068: $$ \frac{1}{\Gamma(\alpha)} F[f^{(\alpha)}](x) \int_0^\infty
1069: \tau^{\alpha-1} e^{-ix\tau}d\tau = (ix)^{-\alpha} F[f^{(\alpha)}](x) =
1070: F[f](x). $$
1071: By Theorem 383 in \cite{hardyetal}, for $p>1$ and
1072: $$ 0 < \alpha < \frac{1}{p}, \qquad q=\frac{p}{1-\alpha p}, $$
1073: $I_\alpha$ maps $L_p$ into $L_q$, and is bounded. That is, there
1074: exists a constant $c(p)>0$ such that
1075: \begin{equation}\label{eq:b7}
1076: ||I_\alpha h||_q \le c(p) ||h||_p.
1077: \end{equation}
1078: Introduce
1079: $\psi(x)=f^{(\alpha)}(x)\mathbf{1}_{(-\infty,\sqrt{2}/2]}(x)$, so that
1080: $\psi$ is supported by $[-\sqrt{2}/2,\sqrt{2}/2]$. According to
1081: \eqref{eq:b6},
1082: $$ (I_\alpha \psi)(x) = f(x), \qquad x\le \sqrt{2}/2. $$
1083: So, using \eqref{eq:b7} and monotonicity of the $L_s$-averages, we
1084: have
1085: $$
1086: ||f||_q \le ||I_\alpha \psi ||_q \le c(p) ||\psi||_p \le c_1(p)
1087: ||\psi||_2 \le c_2(p) ||f^{(\alpha)}||_2, \quad c_1(p) :=
1088: (\sqrt{2})^{1/p-1/2} c(p).
1089: $$
1090: In light of \eqref{eq:b5}, for $\alpha=1/2$ we obtain then
1091: \begin{equation}\label{eq:b8}
1092: ||f||_q^2 \le c_2(p) K(f), \quad c_2(p):= \frac{2}{\pi}c_1(p)^2,
1093: \quad \left(p\in (1,2),\ q=\frac{p}{1-p/2} \right).
1094: \end{equation}
1095: Let $x_0 \in (-\sqrt{2}/2,\sqrt{2}/2)$ be such that
1096: $|f(x_0)|=||f||_\infty$. Since $f$ is $2$-Lipschitz,
1097: $$ |f(x)| \ge ||f||_\infty - 2|x-x_0|, \quad |x-x_0| \le
1098: \frac{||f||_\infty}{2}. $$
1099: Then
1100: $$
1101: ||f||_q^2 \ge \left(2 \int_0^{||f||_\infty/2} (||f||_\infty-2y)^qdy
1102: \right)^{2/q} = \frac{||f||_\infty^{2(q+1)/q}}{(q+1)^{2/q}},
1103: $$
1104: so, using \eqref{eq:b8}, we conclude that, for an absolute constant
1105: $c^*(p,q)>0$,
1106: $$ K(f) \ge c^*(p) ||f||_\infty^{2(q+1)/q}. $$
1107: It remains to observe that
1108: $$ \frac{2(q+1)}{q}=1+\frac{2}{p} $$
1109: can be made arbitrarily close to 2 from above by selecting $p$
1110: sufficiently close to 2 from below. This completes the proof.
1111: \qed
1112:
1113: \paragraph{Theorem 8.} For a Young diagram $\lambda$ whose graph lies
1114: within the $n\times n$ square, let $g_{\lambda}(u)$ be the rotated
1115: coordinate version of the function $f_{\lambda}(x)$ defined in
1116: \eqref{eq:lambdagraph}. Denote $\alpha = k/n^2$.
1117: Then for all $2<r<3$, there are constants $c=c(r)>0,
1118: C=C(r)>0$ such that for any $\epsilon>0$ and for any $n$,
1119: \begin{equation}\label{eq:theorem8}
1120: \pr_n\bigg( T\in{\cal T}_n :
1121: \max_{1\le k\le n^2-1}
1122: ||g_{\lambda_T^{k}}-\tilde{g}_\alpha||_\infty >
1123: \epsilon \bigg) \le C \exp(3n -c \,\epsilon^r n^2).
1124: \end{equation}
1125: Consequently, with probability subexponentially close to $1$, for all
1126: $k$ the supnorm distance between $g_{\lambda_T^k} $ and
1127: $\tilde{g}_{\alpha},\, (\alpha=k/n^2),$ does not exceed
1128: $n^{-1/2+\delta}$, $(\delta>0)$.
1129:
1130: \paragraph{Proof.} Let $p(m)$ be the number of partitions of an
1131: integer $m$. It is known that for all $m$, $p(m) \le \exp(\pi
1132: \sqrt{2m/3})$ (see \cite{apostol}, Theorem 14.5). Fix $n$, $1\le k\le
1133: n^2-1$, $\epsilon>0$. Using \hbox{Lemma 1},
1134: \begin{equation*}
1135: \pr_n\bigg( T\in{\cal T}_n : ||g_{\lambda_T^{k}}
1136: -\tilde{g}_\alpha||_\infty > \epsilon \bigg) =
1137: \sum_{
1138: \begin{array}{ll}\lambda_0\subseteq \square_n\textrm{ of area }k \\
1139: ||g_{\lambda_0}-\tilde{g}_\alpha||_\infty > \epsilon
1140: \end{array}}
1141: \pr_n\bigg( T\in{\cal T}_n : \lambda_T^{k} =
1142: \lambda_0 \bigg)
1143: \end{equation*}
1144: \begin{equation}\label{eq:usinglemma}
1145: \le p(k)\sup_{
1146: \begin{array}{ll}\lambda_0\subseteq\square_n\textrm{ of area }k \\
1147: ||g_{\lambda_0}-\tilde{g}_\alpha||_\infty > \epsilon
1148: \end{array}} \exp\bigg(-(1+o(1))n^2(K(g_{\lambda_0})+H(\alpha)-\log
1149: 2) \bigg).
1150: \end{equation}
1151: Let $\lambda_0$ be a diagram contained in $\square_n$ of area $k$,
1152: such that $||g_{\lambda_0}-\tilde{g}_\alpha||_\infty > \epsilon$.
1153: Since $g_{\lambda_0}$ is $\alpha$-admissible, using Theorem 7 and
1154: Lemma 4 we have
1155: $$ K(g_{\lambda_0})+H(\alpha)-\log 2 \ge
1156: K(g_{\lambda_0}-\tilde{g}_\alpha) >
1157: c(r)||g_{\lambda_0}-\tilde{g}_\alpha||_\infty^r
1158: \ge c(r)\epsilon^r. $$
1159: Combining this with \eqref{eq:usinglemma} and with
1160: the above remark on the number of partitions of an integer gives
1161: that for $n$ larger than some absolute initial bound,
1162: \begin{equation*}
1163: \pr_n\bigg( T\in{\cal T}_n : ||g_{\lambda_T^{k}}
1164: -\tilde{g}_\alpha||_\infty > \epsilon \bigg) \le \exp(2.8 \sqrt{\alpha}n - c n^2
1165: \epsilon^r ).
1166: \end{equation*}
1167: Taking the union bound over all $1\le k\le n^2-1$ gives
1168: \eqref{eq:theorem8}.
1169: \qed
1170:
1171: \paragraph{Lemma 5.} For each $(x,y)\in (0,1)\times(0,1)$, let $(u,v)$
1172: be their rotated coordinates as in \eqref{eq:rotated}. Let $\alpha_0 =
1173: L(x,y)$, so that $|u|<\sqrt{2\alpha_0(1-\alpha_0)}$ and
1174: $v=\tilde{g}_{\alpha_0}(u)$. There exist absolute constants $c_1,
1175: c_2>0$ such that if we set
1176: $$\sigma (x,y)=\min(xy,(1-x)(1-y)),$$
1177: $$d(x,y)=c_1\sqrt{\sigma(x,y)},\quad \Delta(x,y) = c_2 \sigma^2(x,y), $$
1178: we will have that for all $0<\alpha<1$ and $\delta < \Delta(x,y)$, if
1179: $|\tilde{g}_\alpha(u)-\tilde{g}_{\alpha_0}(u)|<\delta\cdot d(x,y)$ then
1180: $|\alpha-\alpha_0| < \delta$.
1181:
1182: \paragraph{Proof.} Since $\tilde{g}_{\alpha}(u)$ increases with $\alpha$, it
1183: suffices to prove existence of two absolute constants $\gamma_1,\gamma_2>0$
1184: such that
1185: $$
1186: |\tilde{g}_{\alpha}(u)-\tilde{g}_{\alpha_0}(u)|\ge \gamma_1\sigma^{1/2}(x,y)
1187: |\alpha-\alpha_0|, \quad\textrm{if }|\alpha-\alpha_0|\le \gamma_2\sigma(x,y).
1188: $$
1189: Because of the symmetry property
1190: $\tilde{g}_{1-\alpha}(u)=\sqrt{2}-\tilde{g}_{\alpha}(u)$, we may
1191: assume that $x+y\le 1$, or equivalently that $\alpha_0\le 1/2$.
1192:
1193: %We claim that for some constants $c_1,c_2>0$, if $\alpha\in(0,1)$
1194: %satisfies $|\alpha-\alpha_0| \le c_2 x y$, then
1195: %$|\tilde{g}_\alpha(u)-\tilde{g}_{\alpha_0}(u)|\ge c_1|\alpha-\alpha_0|
1196: %\sqrt{xy}$. This will prove the statement of the Lemma; for if
1197: %$\alpha\in(0,1)$ and $\delta<c_2 x y$ are such that
1198: %$|\tilde{g}_{\alpha}(u)-\tilde{g}_{\alpha_0}(u)|<\delta\cdot c_1
1199: %\sqrt{xy}$ but $|\alpha-\alpha_0|\ge \delta$, then either
1200: %$|\alpha-\alpha_0|\le c_2 x y$, in which case we get
1201: %$$ c_1 \delta \sqrt{x y} \le c_1 |\alpha_0-\alpha|\sqrt{x y} \le
1202: %|\tilde{g}_\alpha(u)-\tilde{g}_{\alpha_0}(u)| < \delta c_1 \sqrt{x y},
1203: %$$
1204: %a contradiction; or $|\alpha-\alpha_0| \ge c_2 x y$, in which case, if
1205: %$\alpha > \alpha_0 + c_2 x y$ we get
1206: %\begin{eqnarray*}
1207: %c_1 c_2 xy \sqrt{x y} &>&
1208: %|\tilde{g}_\alpha(u)-\tilde{g}_{\alpha_0}(u)| \ge
1209: %|\tilde{g}_{\alpha_0+c_2 x y}(u)-\tilde{g}_{\alpha_0}(u)| \\ &\ge&
1210: %c_1 |(\alpha_0+c_2 x y) - \alpha_0| \sqrt{x y} = c_1 c_2 x y\sqrt{x
1211: %y},
1212: %\end{eqnarray*}
1213: %again a contradiction, or a similar contradiction if $\alpha <
1214: %\alpha_0 - c_2 x y$.
1215:
1216: To prove the above claim, we note the following inequalities. Notice
1217: first that
1218: $$ \sqrt{2\alpha_0(1-\alpha_0)} \ge v \implies \alpha_0 \ge \frac{1-\sqrt{1-2
1219: v^2}}{2}. $$
1220: Likewise, $\alpha^{(-)}$ that corresponds to the lowest point $(u,u)$ is given
1221: by
1222: $$ \alpha^{(-)} = \frac{1-\sqrt{1-2 u^2}}{2}. $$
1223: and we see that
1224: \begin{equation}\label{eq:lowest}
1225: \alpha_0 - \alpha^{(-)} \ge \frac{\sqrt{1-2 u^2}-\sqrt{1-2 v^2}}{2}
1226: = \frac{v^2-u^2}{\sqrt{1-2u^2}+\sqrt{1-2 v^2}} \ge \frac{v^2-u^2}{2}
1227: = xy.
1228: \end{equation}
1229: \eqref{eq:lowest} says that decreasing $\alpha_0$ by $x_0 y_0$ gives
1230: us a feasible $\alpha$, for which $(u,\tilde{g}_\alpha(u))$ lies
1231: between $(u,v)$ and the lowest point $(u,u)$, such that $u\le\sqrt{2\alpha
1232: (1-\alpha)}$.
1233:
1234: Let us estimate from above $\tilde{g}_\alpha(u)$ for
1235: $\alpha\in[\alpha^{(-)},\alpha_0]$. From \eqref{eq:partialdiff} it
1236: follows that
1237: $$ \frac{\partial \tilde{g}_\alpha(u)/\partial
1238: \alpha}{\sqrt{\tilde{g}_\alpha(u)^2-u^2}} \ge c $$
1239: for some absolute constant $c>0$. (Indeed,
1240: $2\alpha(1-\alpha)=\beta^2(\alpha) \ge \tilde{g}_\alpha(u)^2$.)
1241: Integrating from $\alpha\in [\alpha^{(-)},\alpha_0]$ and
1242: exponentiating, we obtain
1243: $$ \frac{\tilde{g}_{\alpha_0}(u) + \sqrt{\tilde{g}_{\alpha_0}(u)^2-u^2}}
1244: {\tilde{g}_{\alpha}(u) + \sqrt{\tilde{g}_{\alpha}(u)^2-u^2}} \ge
1245: \exp(c(\alpha_0-\alpha)),$$
1246: or equivalently
1247: $$ \frac{\tilde{g}_{\alpha}(u) - \sqrt{\tilde{g}_{\alpha}(u)^2-u^2}}
1248: {\tilde{g}_{\alpha_0}(u) - \sqrt{\tilde{g}_{\alpha_0}(u)^2-u^2}} \ge
1249: \exp(c(\alpha_0-\alpha)).$$
1250: Consequently
1251: $$ \sqrt{\tilde{g}_\alpha(u)^2-\tilde{g}_{\alpha_0}(u)^2} \le
1252: \cosh(c(\alpha_0-\alpha)) \sqrt{\tilde{g}_{\alpha_0}(u)^2-u^2} -
1253: \sinh(c(\alpha_0-\alpha))\tilde{g}_{\alpha_0}(u), $$
1254: or
1255: $$ \tilde{g}_\alpha(u)^2 \le \bigg[\cosh(c(\alpha_0-\alpha))
1256: \tilde{g}_{\alpha_0}(u) - \sinh(c(\alpha-\alpha_0))
1257: \sqrt{\tilde{g}_{\alpha_0}(u)^2 - u^2}\bigg]^2, $$
1258: so that
1259: $$ \tilde{g}_{\alpha}(u) \le
1260: \cosh(c(\alpha_0-\alpha))
1261: \tilde{g}_{\alpha_0}(u) - \sinh(c(\alpha-\alpha_0))
1262: \sqrt{\tilde{g}_{\alpha_0}(u)^2 - u^2}. $$
1263: Consequently, for some constants $c_i > 0$,
1264: \begin{eqnarray*}
1265: \tilde{g}_{\alpha}(u) - \tilde{g}_{\alpha_0}(u) &\le& -c_3
1266: (\alpha_0-\alpha) [ (v^2-u^2)^{1/2} - c_4 (\alpha_0-\alpha)v ] \\
1267: &=& - c_5(\alpha_0-\alpha)[ (x y)^{1/2} - c_6 (\alpha_0-\alpha)(x+y)]
1268: \\ &\le& - c_7 (\alpha_0-\alpha) (x y)^{1/2},
1269: \end{eqnarray*}
1270: provided that
1271: $$ \alpha_0-\alpha \le c_8 \frac{(x y)^{1/2}}{x+y}. $$
1272: From \eqref{eq:lowest} we know that we can go below $\alpha_0$ by $xy$ at
1273: least.
1274: Pick $\rho = \min(1,c_8/3)$; then the last inequality
1275: holds for $\alpha_0-\alpha\le \rho x y$, and we have
1276: \begin{equation}\label{eq: smallalpha}
1277: \tilde{g}_{\alpha}(u) - \tilde{g}_{\alpha_0}(u) \le
1278: - c_7 (\alpha_0-\alpha)(x y)^{1/2}, \qquad \alpha\in [\alpha_0 - \rho
1279: x y, \alpha_0].
1280: \end{equation}
1281: Now for $\alpha_0\le \alpha\le 1/2$ we know that
1282: $$ \frac{\partial \tilde{g}_{\alpha}(u)}{\partial \alpha} \ge c_9
1283: \sqrt{v^2-u^2} = c_{10} (x y)^{1/2}, $$
1284: so that
1285: \begin{equation}\label{eq: moderatealpha1}
1286: \tilde{g}_\alpha(u) - \tilde{g}_{\alpha_0}(u) \ge c_{10} (x
1287: y)^{1/2} (\alpha-\alpha_0).
1288: \end{equation}
1289: By symmetry, for $1/2\le\alpha\le 1-\alpha_0$,
1290: \begin{equation}\label{eq: moderatealpha2}
1291: \tilde{g}_{\alpha}(u)-\tilde{g}_{1-\alpha_0}\le -c_{10}((1-\alpha_0)-\alpha)
1292: ((1-x)(1-y))^{1/2},
1293: \end{equation}
1294: and, for $1-\alpha_0\le\alpha\le 1-\alpha_0+\rho(1-x)(1-y)$,
1295: \begin{equation}\label{eq: largealpha}
1296: \tilde{g}_{\alpha}(u)-\tilde{g}_{\alpha_0}(u)\ge c_7(\alpha-(1-\alpha_0))
1297: ((1-x)(1-y))^{1/2}.
1298: \end{equation}
1299: The inequalities \eqref{eq: smallalpha}, \eqref{eq: moderatealpha1},
1300: \eqref{eq: moderatealpha2}, \eqref{eq: largealpha} prove the claim with
1301: $\gamma_1=\min\{c_7,c_{10}\}$ and $\gamma_2=\rho$.
1302: \qed
1303:
1304: \paragraph{Proof of Theorem 1(ii).} We now prove Theorem 1(ii), the
1305: part of Theorem 1 that deals with the interior of the square. The
1306: treatment of the boundary of the square is more delicate and is
1307: deferred to section 4, being essentially equivalent to Theorem 3.
1308:
1309: Fix $1\le i, j\le n$ such that
1310: \begin{equation}\label{eq:interior}
1311: \min(i j,(n-i)(n-j)) > n^{3/2+\epsilon}.
1312: \end{equation}
1313: Let $(u,v)$ be the rotated coordinates
1314: corresponding to $(x,y)=(i/n,j/n)$. Let $\alpha_0 = L(i/n,j/n)$, so
1315: that $v=\tilde{g}_{\alpha_0}(u)$ and
1316: $|u|<\sqrt{2\alpha_0(1-\alpha_0)}$. For each tableau
1317: $T=(t_{i,j})_{1\le i,j\le n}\in{\cal T}_n$ let $k_T = t_{i,j}$, and let
1318: $\beta_T = k_T/n^2$. Note that $k_T$ is an integer representing the
1319: smallest $s$ such that $\lambda_T^s$ contains the box $(i,j)$. This
1320: implies that
1321: $$ v \le g_{\lambda_T^{k_T}}(u) \le v+\frac{\sqrt{2}}{n}, $$ or
1322: \begin{equation}\label{eq:sqrtn}
1323: \left| g_{\lambda_T^{k_T}}(u)-\tilde{g}_{\alpha_0}(u) \right| \le
1324: \frac{\sqrt{2}}{n}.
1325: \end{equation}
1326: Apply Lemma 5 with $(x,y)=(i/n,j/n)$ and
1327: $\delta=n^{-(1-\epsilon)/2}$. Note that because of
1328: \eqref{eq:interior}, for $n$ large we have $\delta < \Delta(x,y)$ as
1329: required. Then, making use of Theorem 8 we get
1330: \begin{eqnarray*}
1331: \pr_n\left( T\in{\cal T}_n : \left| \frac{1}{n^2}t_{i,j}-
1332: L\left(\frac{i}{n},\frac{j}{n}\right)
1333: \right| > \frac{1}{n^{(1-\epsilon)/2}} \right)
1334: &=&
1335: \pr_n\Big( T\in {\cal T}_n : |\beta_T - \alpha_0 | >
1336: \frac{1}{n^{(1-\epsilon)/2}} \Big)
1337: \end{eqnarray*}
1338:
1339: \vspace{-20.0pt}
1340: \begin{eqnarray*}
1341: \textrm{(by Lemma 5)}
1342: &\le& \pr_n\Big( T\in {\cal T}_n :
1343: |\tilde{g}_{\beta_T}(u)-\tilde{g}_{\alpha_0}(u) | >
1344: \frac{d(i/n,j/n)}{n^{(1-\epsilon)/2}} \Big) \\
1345: \textrm{(by \eqref{eq:sqrtn})}
1346: &\le& \pr_n\Big( T\in{\cal T}_n : \left| g_{\lambda_T^{k_T
1347: }}(u) - \tilde{g}_{\beta_T}(u) \right| >
1348: \frac{d(i/n,j/n)}{n^{(1-\epsilon)/2}}-\frac{\sqrt{2}}{n} \Big)
1349: \\ \textrm{(for $n$ large enough, by \eqref{eq:interior})} &\le&
1350: \pr_n\Big( T\in{\cal T}_n : \left| g_{\lambda_T^{k_T
1351: }}(u) - \tilde{g}_{\beta_T}(u) \right| >
1352: \frac{d(i/n,j/n)}{2 n^{(1-\epsilon)/2}} \Big)
1353: \\ \textrm{(by Theorem 8 with $r=2+\epsilon$)}
1354: &\le& C \exp\left(3n - c n^2 \left(\frac{d(i/n,j/n)}{2
1355: n^{(1-\epsilon)/2}}\right)^{2+\epsilon} \right) \\
1356: \textrm{(for $n$ large, by \eqref{eq:interior})}
1357: & \le & C' \exp(-c' n^{3/2}).
1358: \end{eqnarray*}
1359: Taking the union bound over all $1\le i,j\le n$ satisfying
1360: \eqref{eq:interior} gives the result.
1361: \qed
1362:
1363: \section{Solution of the variational problem}
1364:
1365: \subsection{Preliminaries}
1366:
1367: In this section, we prove Theorem 7. We actually \emph{derive} the
1368: explicit formula for the minimizer using methods of the calculus of
1369: variations and the theory of singular (Cauchy-type) integral
1370: equations. Our derivation makes only one a priori assumption (obtained
1371: by educated guesswork and later verified by computer simulations) on
1372: the graphical form that the minimizer would take, and so is in a sense
1373: more systematic than the analogous treatments in the fundamental
1374: papers \cite{loganshepp}, \cite{vershikkerov1}, \cite{vershikkerov2},
1375: where the solutions are brilliantly guessed using the properties of
1376: the Hilbert transform. We believe that our technique may prove useful
1377: in the treatment of similar problems in the future.
1378:
1379: First, observe that because of symmetry, we need only treat the
1380: case $\alpha \le 1/2$; the mapping $g \to \sqrt{2}-g$ takes the set of
1381: $\alpha$-admissible functions bijectively onto the set of
1382: $(1-\alpha)$-admissible functions, and has the property that
1383: $K(\sqrt{2}-g)=K(g)$.
1384:
1385: Next, observe that for $\alpha=1/2$ the assertion is immediate,
1386: because of Lemma 3.
1387:
1388: We prove another fact that follows from general considerations, before
1389: turning to the derivation of the minimizer.
1390:
1391: \paragraph{Lemma 6.} For any $0<\alpha<1$, the functional $K$ has a unique
1392: $\alpha$-admissible minimizer.
1393:
1394: \paragraph{Proof.} The functional $K$ is continuous on the space of
1395: $\alpha$-admissible functions, and is bounded below by Lemma 3. By the
1396: Arzela-Ascoli theorem, the space of $\alpha$-admissible functions is
1397: compact in the topology induced by the supremum norm (since the
1398: admissible functions are uniformly bounded and
1399: equicontinuous). Therefore $K$ has a minimizer. To prove that the
1400: minimizer is unique, let $h_1$ and $h_2$ be two distinct
1401: $\alpha$-admissible minimizers. Then $\tilde{h}=(h_1+h_2)/2$ is also
1402: an $\alpha$-admissible function, and $g=(h_1-h_2)/2 \nequiv 0$, $g(\pm
1403: \sqrt{2}/2) = 0$. So, using the parallelogram identity and Lemma 3,
1404: $$ K(\tilde{h}) = \frac{1}{2}K(h_1)+\frac{1}{2}K(h_2)-K(g) < \min_{h
1405: \textrm{ is }\alpha\textrm{-admissible}} K(h), $$
1406: a contradiction.
1407: \qed
1408:
1409: \subsection{The derivation}
1410:
1411: We now proceed with the derivation of the minimizer, which we shall
1412: denote $h=h_\alpha$. The dependence on $\alpha$ will be suppressed
1413: except where it is required. For the rest of this section, $\alpha$
1414: will be a fixed value in $(0,1/2)$, unless stated otherwise.
1415:
1416: First, note that, under the condition $h(\pm\sqrt{2}/2)=\sqrt{2}/2$,
1417: the $\alpha$-condition $\int_{-\sqrt{2}/2}^{\sqrt{2}/2}
1418: (h(u)-|u|)du=\alpha$ is equivalent to
1419: \begin{equation}\label{eq:alphacondition}
1420: -\int_{-\sqrt{2}/2}^{\sqrt{2}/2} u h'(u) du = \alpha - \frac{1}{2}.
1421: \end{equation}
1422: %We now formulate a sufficient condition for $h$ to be a
1423: %minimizer. This condition is actually also necessary, and knowing this
1424: %is essential to understanding why the derivation works. However,
1425: %since we are ultimately interested mainly in proving Theorem 7 (the
1426: %``verification'' part), we omit the proof of necessity, which consists
1427: %of roughly the same ideas with some additional logical contortions.
1428: %
1429: We now formulate a sufficient condition for $h$ to be a minimizer. It
1430: is based on a standard recipe of the calculus of variations, the
1431: Lagrange formalism. We form the Lagrange function
1432: $$ {\cal L}(h,\lambda) = K(h) - \lambda
1433: \int_{-\sqrt{2}/2}^{\sqrt{2}/2} u h'(u) du $$
1434: and require that, for some $\lambda, h_\alpha$ be a local minimum
1435: point of ${\cal L}(h,\lambda)$ in the convex set of functions $h$
1436: subject to all the restrictions except the $\alpha$-condition
1437: \eqref{eq:alphacondition}. To be sure, we ought to include into the
1438: function a term $\lambda'$ times the integral of $h'$, since $h$ must
1439: meet another constraint
1440: \begin{equation}\label{eq:integral0}
1441: \int_{-\sqrt{2}/2}^{\sqrt{2}/2} h'(u)du = 0.
1442: \end{equation}
1443: We chose not to, since -- in the square case -- even without this
1444: constraint $h'(u)$ will turn out to be odd anyway. Since ${\cal
1445: L}(h,\lambda)$ depends explicitly on $h'$ alone, we get the equations
1446: for the sufficient condition in a simple-minded manner, by taking
1447: partial derivatives of ${\cal L}$ with respect to $h'(s),\ s\in
1448: (-\sqrt{2}/2,\sqrt{2}/2)$, and paying attention only to the constraint
1449: $-1 \le h'(s) \le 1$. The resulting ``complementary slackness''
1450: conditions are
1451: \begin{equation}\label{eq:necessarycond}
1452: w(s):= - \int_{-\sqrt{2}/2}^{\sqrt{2}/2} h'(t)\log|s-t|dt - \lambda s
1453: \ \ \ \textrm{ is } \left\{
1454: \begin{array}{ll}
1455: = 0, & \textrm{if }-1 < h'(s) < 1, \\
1456: \ge 0, & \textrm{if }h'(s)=-1, \\
1457: \le 0, & \textrm{if }h'(s)=1.
1458: \end{array}
1459: \right.
1460: \end{equation}
1461:
1462: \paragraph{Lemma 7.} If $h$ is an $\alpha$-admissible function that,
1463: for some $\lambda\in\mathbb{R}$, satisfies \eqref{eq:necessarycond}
1464: for all $s\in(-\sqrt{2}/2,\sqrt{2}/2)$ for which $h'(s)$ is defined,
1465: then $h$ is a minimizer.
1466:
1467: \paragraph{Proof.} If $g$ is a $1$-Lipschitz function on
1468: $[-\sqrt{2}/2,\sqrt{2}/2]$, then \eqref{eq:necessarycond} implies
1469: that $(g'(s)-h'(s))w(s) \ge 0$ for all $s$ for which this is defined,
1470: so
1471: $$
1472: \int_{-\sqrt{2}/2}^{\sqrt{2}/2} g'(s)w(s)ds \ge
1473: \int_{-\sqrt{2}/2}^{\sqrt{2}/2} h'(s)w(s)ds. $$
1474: If $g$ is $\alpha$-admissible, by \eqref{eq:alphacondition} this can
1475: be written as
1476: \begin{eqnarray*} 2 \langle h,g \rangle + \alpha - \frac{1}{2} &=&
1477: 2 \langle h,g \rangle - \lambda \int_{-\sqrt{2}/2}^{\sqrt{2}/2} s
1478: g'(s) ds \\ &\ge& 2 \langle h,h \rangle - \lambda
1479: \int_{-\sqrt{2}/2}^{\sqrt{2}/2} s h'(s)ds = 2 \langle h,h \rangle +
1480: \alpha - \frac{1}{2},
1481: \end{eqnarray*}
1482: which shows that
1483: $$ \langle h,g\rangle \ge \langle h,h \rangle. $$
1484: Therefore, by Lemma 3 applied to the function $g-h$,
1485: $$
1486: \langle g,g \rangle = \langle h,h \rangle + 2 \langle h, g-h \rangle +
1487: \langle g-h, g-h \rangle \ge \langle h,h\rangle,
1488: $$
1489: so $h$ is a minimizer. \qed
1490:
1491: \bigskip
1492: We are about to prove part (i) of Theorem 7, namely that
1493: $h=\tilde{g}_\alpha$ is the minimizer. Assuming this, note that in the
1494: above proof we actually showed that
1495: $$ \langle g,g \rangle \ge \langle \tilde{g}_\alpha, \tilde{g}_\alpha
1496: \rangle + \langle g-\tilde{g}_\alpha, g-\tilde{g}_\alpha \rangle, $$
1497: which is precisely the claim of part (iii) of Theorem 7. So it remains
1498: to prove parts (i) and (ii).
1499:
1500: Our challenge now is to determine an admissible $h$ that meets the
1501: conditions \eqref{eq:necessarycond}. Now look at Figure 1(c) with your
1502: head tilted 45 degrees to the right. Based on the shape of the level
1503: curves, we make the following assumption: For some
1504: $\beta=\beta(\alpha) \in (0,\sqrt{2}/2)$,
1505: \begin{equation}\label{eq:guess}
1506: h'(s)\ \ \ \textrm{ is }\left\{ \begin{array}{ll}
1507: = -1, & \textrm{if }-\sqrt{2}/2 < s < -\beta, \\
1508: \in (-1,1), & \textrm{if }-\beta<s<\beta, \\
1509: = +1, & \textrm{if }\beta<s<\sqrt{2}/2. \end{array}\right.
1510: \end{equation}
1511: Substituting this into \eqref{eq:necessarycond} gives that for
1512: $-\beta<s<\beta$,
1513: \begin{multline}\label{eq:multline}
1514: -\int_{-\beta}^\beta h'(t)\log|s-t|dt = \lambda s -
1515: \int_{-\sqrt{2}/2}^\beta \log(s-t)dt + \int_\beta^{\sqrt{2}/2}
1516: \log(t-s)dt \\ =\lambda s +
1517: (\sqrt{2}/2-s)\log(\sqrt{2}/2-s)-(\sqrt{2}/2+s)\log(\sqrt{2}/2+s)
1518: \\ + (\beta+s)\log(\beta+s) - (\beta-s)\log(\beta-s)
1519: \qquad\qquad\qquad\qquad\qquad
1520: \end{multline}
1521: Assume that $h'(s)$ is continuously differentiable on
1522: $(-\beta,\beta)$. Differentiate \eqref{eq:multline}, to obtain
1523: \begin{equation}\label{eq:diffmultline}
1524: -\int_{-\beta}^\beta \frac{h'(t)}{s-t}dt = \lambda+ \log
1525: \frac{\beta^2-s^2}{\frac{1}{2}-s^2},
1526: \end{equation}
1527: where the left-hand side is a principal value integral.
1528:
1529: In the theory of integral equations this is known as an airfoil
1530: equation. Solving it is tantamount to inverting a Hilbert transform on
1531: a finite interval. Fortunately for us, it can be solved! The following
1532: theorem appears in \cite{estradakanwal}, Section 3.2, p. 74. (See also
1533: \cite{porterstirling}, Section 9.5.2.)
1534:
1535: \paragraph{Theorem 9.}
1536: The general solution of the airfoil equation
1537: $$
1538: \frac{1}{\pi}\int_{-1}^1\frac{g(y)}{y-x}\,dx=f(x),\quad |x|<1,
1539: $$
1540: with the integral understood in the principal value sense, and $f(x)$
1541: satisfying a H\"older condition, is given by
1542: \begin{equation}\label{eq:airfoil}
1543: g(x)=\frac{1}{\pi\sqrt{1-x^2}}\int_{-1}^1\frac{\sqrt{1-y^2}f(y)}{x-y}\,dy
1544: +\frac{c}{\sqrt{1-x^2}}.
1545: \end{equation}
1546:
1547: \bigskip
1548: Applying Theorem 9 to \eqref{eq:diffmultline}, we get the equation
1549: \begin{equation}\label{eq:kanwal}
1550: h'(s) = \frac{1}{\pi^2(\beta^2-s^2)^{1/2}} \int_{-\beta}^\beta
1551: (\beta^2-t^2)^{1/2} \left( \lambda+\log
1552: \frac{\beta^2-t^2}{\frac{1}{2}-t^2} \right) \frac{dt}{s-t}
1553: + \frac{c}{(\beta^2-s^2)^{1/2}}.
1554: \end{equation}
1555: Here the integral is again in the sense of principal value, and the
1556: equation must hold for some value of $c$.
1557:
1558: We evaluate the integral in \eqref{eq:kanwal}. Consider the
1559: contribution of the $\lambda$-term first. Substituting $t=\beta \sin
1560: x$ and later $u=\tan x/2$, we get
1561: \begin{eqnarray*}
1562: \int_{-\beta}^\beta \frac{(\beta^2-t^2)^{1/2}}{s-t}dt &=& \beta
1563: \int_{-\pi/2}^{\pi/2} \frac{\cos^2 x}{s/\beta - \sin x}dx
1564: \\ &=& \beta
1565: \int_{-\pi/2}^{\pi/2} (s/\beta+\sin x)dx + \beta \left(1-(s/\beta)^2
1566: \right) \int_{-\pi/2}^{\pi/2} \frac{dx}{s/\beta-\sin x}
1567: \\ &=& \pi s + \frac{2\beta\left(1-(s/\beta)^2\right)}{s/\beta}
1568: \int_{-1}^1 \frac{du}{u^2 - 2(\beta/s)u + 1}.
1569: \end{eqnarray*}
1570: For $|s|<\beta$, the denominator in the last integral has two real
1571: roots, $u_1\in(-1,1)$ and $u_2\notin(-1,1)$. A simple computation
1572: shows that the principal value of this integral at $u=u_1$ is zero. So
1573: \begin{equation}\label{eq:lambdaterm}
1574: \int_{-\beta}^\beta \frac{(\beta^2-t^2)^{1/2}}{s-t}dt = \pi s, \qquad
1575: s\in(-\beta,\beta).
1576: \end{equation}
1577: Turn to the log-part of the integral in
1578: \eqref{eq:kanwal}. Substituting $t=\tau \beta$, $s= v_1 \beta$,
1579: $(2\beta^2)^{-1} = v_2^2$, we see that
1580: \begin{equation}\label{eq:logterm}
1581: \int_{-\beta}^\beta \frac{(\beta^2-t^2)^{1/2}}{s-t} \log
1582: \frac{\beta^2-t^2}{\frac{1}{2}-t^2} dt = \beta [
1583: I(s/\beta,\sqrt{2}/(2\beta)) - I(-s/\beta,\sqrt{2}/(2\beta)) ],
1584: \end{equation}
1585: where
1586: $$ I(\xi,\gamma) = \int_{-1}^1 \frac{(1-\eta)^{1/2}}{\xi-\eta}
1587: \log\frac{1+\eta}{\gamma+\eta} d\eta, \qquad \xi\in[-1,1], \gamma\ge 1,
1588: $$
1589: is evaluated in the following lemma.
1590:
1591: \paragraph{Lemma 8.}
1592: \begin{multline}\label{eq:evaluation}
1593: I(\xi,\gamma) = \pi\bigg[ 1-\gamma+\sqrt{\gamma^2-1} -
1594: \xi\log\left(\gamma+\sqrt{\gamma^2-1}\right) \\
1595: - 2\sqrt{1-\xi^2} \arct \sqrt{
1596: \frac{(\gamma-1)(1-\xi)}{(\gamma+1)(1+\xi)} } \ \bigg].
1597: \qquad\qquad\qquad\qquad
1598: \end{multline}
1599:
1600: \paragraph{Proof.} Notice that $I(\xi,1)=0$, and, for $x>1$,
1601: \begin{multline}\label{eq:multline2}
1602: \frac{\partial I(\xi,x)}{\partial x} = - \int_{-1}^1
1603: \frac{(1-\eta^2)^{1/2}}{(\xi-\eta)(x+\eta)}d\eta \\
1604: = -\frac{1}{x+\xi}\left[ \int_{-1}^1
1605: \frac{(1-\eta^2)^{1/2}}{\xi-\eta}d\eta + \int_{-1}^1
1606: \frac{(1-\eta^2)^{1/2}}{x+\eta}d\eta \right] \qquad\qquad\qquad\quad\\
1607: = -\frac{\pi\xi}{x+\xi}-\frac{1}{x+\xi}\int_{-1}^1
1608: \frac{(1-\eta^2)^{1/2}}{x+\eta}d\eta,
1609: \qquad\qquad\qquad\qquad\qquad\qquad\qquad\ \ \,
1610: \end{multline}
1611: see \eqref{eq:lambdaterm}. Substituting $\eta=\sin t$, $(t\in
1612: [-\pi/2,\pi/2])$, and then $t=2 \arct u$, $(u\in[-1,1])$, we evaluate
1613: \begin{multline}\label{eq:unnumbered}
1614: \int_{-1}^1 \frac{(1-\eta^2)^{1/2}}{x+\eta}d\eta = (xt+\cos
1615: t)|_{\pi/2}^{\pi/2} +(1-x^2) \int_{-\pi/2}^{\pi/2} \frac{dt}{x+\sin t}
1616: \\ = \pi x + 2(1-x^2)\int_{-1}^1 \frac{du}{x(1+u^2)+2u} =
1617: \qquad\ \ \ \ \,\\
1618: \qquad\quad
1619: = \pi x - 2(x^2-1)^{1/2} \left[ \arct \sqrt{\frac{x+1}{x-1}} +
1620: \arct \sqrt{\frac{x-1}{x+1}} \right] \\
1621: = \pi (x-(x^2-1)^{1/2}).
1622: \qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\quad\ \,\,
1623: \end{multline}
1624: Combining this with \eqref{eq:multline2}, we obtain
1625: $$
1626: \frac{\partial I(\xi,x)}{\partial x} = -\pi +
1627: \frac{\pi(x^2-1)^{1/2}}{x+\xi}.
1628: $$
1629: We integrate this equation from $x=1$ to $x=\gamma>1$, and use the
1630: substitutions $x=\cosh t$, $t\in[0,t_0]$, with
1631: $$ t_0 = \textrm{arccosh}\, \gamma = \log\left(
1632: \gamma+(\gamma^2-1)^{1/2} \right), $$
1633: and then $u=e^t$, $u\in[1,u_0]$, with
1634: $$ u_0 = e^{t_0} = \gamma + (\gamma^2-1)^{1/2}. $$
1635: We have
1636: \begin{multline}\label{eq:multline3}
1637: I(\xi,\gamma) = -\pi(\gamma-1) + \pi \int_0^{t_0} \frac{\sinh^2
1638: t}{\cosh t + \xi}dt \\
1639: = - \pi(\gamma-1)+\pi \left[ (\sinh t-\xi t)|_0^{t_0} + 2(\xi^2-1)
1640: \int_0^{u_0} \frac{du}{u^2+2\xi u +1} \right]. \qquad\quad\ \ \,
1641: \end{multline}
1642: Here
1643: \begin{equation}\label{eq:multline35}
1644: (\sinh t-\xi t)|_0^{t_0} = (\gamma^2-1)^{1/2}-\xi
1645: \log\left(\gamma+(\gamma^2-1)^{1/2} \right),
1646: \end{equation}
1647: and the last integral equals
1648: \begin{multline}\label{eq:multline4}
1649: \frac{1}{\sqrt{1-\xi^2}}\arct \frac{u+\xi}{(1-\xi^2)^{1/2}}
1650: \bigg|_1^{u_0} = \frac{1}{\sqrt{1-\xi^2}} \arct
1651: \frac{(u_0-1)(1-\xi^2)^{1/2}}{1-\xi^2+(u_0+\xi)(1+\xi)} \\
1652: = \frac{1}{\sqrt{1-\xi^2}}
1653: \arct\frac{u_0-1}{u_0+1}\sqrt{\frac{1+\xi}{1-\xi}} =
1654: \frac{1}{\sqrt{1-\xi^2}} \arct
1655: \sqrt{\frac{(\gamma-1)(1-\xi)}{(\gamma+1)(1+\xi)}}.
1656: \end{multline}
1657: Combining \eqref{eq:multline3}, \eqref{eq:multline35},
1658: \eqref{eq:multline4} gives \eqref{eq:evaluation}. \qed
1659:
1660: \bigskip
1661: Now from \eqref{eq:lambdaterm}, \eqref{eq:logterm} and
1662: \eqref{eq:evaluation} we get
1663: \begin{multline}\label{eq:arctrad}
1664: \qquad\qquad
1665: h'(s) = \frac{c}{(\beta^2-s^2)^{1/2}} +
1666: \frac{s}{\pi(\beta^2-s^2)^{1/2}}\left( \lambda- 2 \log
1667: \frac{1+\sqrt{1-2 \beta^2}}{\sqrt{2} \beta} \right) \\
1668: + \frac{2}{\pi} \left(
1669: \arct \sqrt{\frac{(\gamma-1)(1+\xi)}{(\gamma+1)(1-\xi)}} -
1670: \arct \sqrt{\frac{(\gamma-1)(1-\xi)}{(\gamma+1)(1+\xi)}} \right),
1671: \qquad
1672: \end{multline}
1673: with $\xi=s/\beta$, $\gamma=\sqrt{2}/(2\beta)$, or, after some
1674: simplification,
1675: \begin{eqnarray*}
1676: h'(s) &=& \frac{c}{(\beta^2-s^2)^{1/2}} +
1677: \frac{s}{\pi(\beta^s-s^2)^{1/2}}\left( \lambda- 2 \log
1678: \frac{1+\sqrt{1-2 \beta^2}}{\sqrt{2} \beta} \right) \\
1679: & & + \frac{2}{\pi} \arct\frac{(1-2\beta^2)^{1/2}
1680: s}{(\beta^2-s^2)^{1/2}}\,.
1681: \end{eqnarray*}
1682: We now observe that the only values of $c$ and $\lambda$ for which
1683: the right-hand side is bounded as $s\nearrow \beta$, $s\searrow-\beta$,
1684: and therefore has a chance of being the derivative of an
1685: $\alpha$-admissible function, are
1686: \begin{equation}\label{eq:lambdac}
1687: c = 0, \qquad \lambda= 2 \log\frac{1+\sqrt{1-2 \beta^2}}{\sqrt{2}
1688: \beta}.
1689: \end{equation}
1690: Therefore we get
1691: \begin{equation}\label{eq:hprime}
1692: h'(s) = \frac{2}{\pi} \arct\frac{(1-2\beta^2)^{1/2}
1693: s}{(\beta^2-s^2)^{1/2}}.
1694: \end{equation}
1695: Note that $h'(s) \in (-1,1)$. We have determined $h'(s)$, except the
1696: value of $\beta=\beta(\alpha)$ such that $h$ is $\alpha$-admissible,
1697: i.e., satisfies \eqref{eq:alphacondition}. Rewrite
1698: \eqref{eq:alphacondition} as
1699: \begin{equation}\label{eq:alphabetacond}
1700: \int_{-\beta}^\beta s h'(s)ds = \alpha - \beta^2.
1701: \end{equation}
1702: Besides evaluating this last integral, to compute $h(s)$ explicitly we
1703: will need $\int_{-\beta}^s h'(u)du$. To this end, integrating the
1704: first arctangent-of-radical function in \eqref{eq:arctrad} on the
1705: interval $[-1,\xi]$, $(\xi\in(-1,1])$, we get
1706: \begin{multline} \label{eq:eqx1}
1707: \qquad
1708: \int_{-1}^\xi \arct
1709: \sqrt{\frac{(\gamma-1)(1+\eta)}{(\gamma+1)(1-\eta)}} d\eta \\ =
1710: \xi \arct \sqrt{\frac{(\gamma-1)(1+\eta)}{(\gamma+1)(1-\eta)}} -
1711: \frac{\sqrt{\gamma^2-1}}{2} \int_{-1}^\xi \frac{\eta\,
1712: d\eta}{(\gamma-\eta)\sqrt{1-\eta^2}}. \qquad
1713: \end{multline}
1714: Substituting in the last integral $\eta=\sin t$, and then $u=\tan t$,
1715: we transform it into
1716: \begin{multline}\label{eq:eqx2}
1717: \qquad
1718: -t_0 - \frac{\pi}{2} + \gamma \int_{-\pi/2}^{t_0}
1719: \frac{dt}{\gamma-\sin t} \qquad
1720: \qquad \qquad\qquad [t_0 = \arcs \xi] \\
1721: = - t_0 - \frac{\pi}{2} + 2 \int_{-1}^{u_0}
1722: \frac{du}{1+u^2-2u/\gamma} \qquad \quad [u_0 = \tan t_0/2]
1723: \qquad\qquad \\
1724: = -t_0 - \frac{\pi}{2} + \frac{2\gamma}{\sqrt{\gamma^2-1}}
1725: \left( \arct \frac{u_0-\gamma^{-1}}{\sqrt{1-\gamma^{-2}}}+
1726: \arct \frac{1+\gamma^{-1}}{\sqrt{1-\gamma^{-2}}} \right)
1727: \ \ \
1728: \\ =
1729: -t_0 - \frac{\pi}{2} + \frac{2\gamma}{\sqrt{\gamma^2-1}}\arct \left(
1730: \frac{1+u_0}{1-u_0} \sqrt{\frac{\gamma-1}{\gamma+1}} \right);
1731: \qquad \qquad \qquad\qquad\quad\ \
1732: \end{multline}
1733: here
1734: \begin{equation}\label{eq:eqx3}
1735: \frac{1+u_0}{1-u_0} = \frac{1+\tan t_0/2}{1-\tan t_0/2} = \frac{1+\sin
1736: t_0}{\cos t_0} = \frac{1+\xi}{\sqrt{1-\xi^2}} =
1737: \sqrt{\frac{1+\xi}{1-\xi}}.
1738: \end{equation}
1739: From \eqref{eq:eqx1}, \eqref{eq:eqx2}, \eqref{eq:eqx3} we obtain
1740: \begin{multline}\label{eq:eqx4} \qquad
1741: \int_{-1}^\xi \arct
1742: \sqrt{\frac{(\gamma-1)(1+\eta)}{(\gamma+1)(1-\eta)}}d\eta \\
1743: = (\xi-\gamma) \arct
1744: \sqrt{\frac{(\gamma-1)(1+\eta)}{(\gamma+1)(1-\eta)}} +
1745: \frac{\sqrt{\gamma^2-1}}{2}\left(\arcs\xi+\frac{\pi}{2}\right). \qquad
1746: \end{multline}
1747: In a similar fashion
1748: \begin{equation}\label{eq:eqx5}
1749: \int_{-1}^1 \eta \arct \sqrt{
1750: \frac{(1+\eta)(\gamma-1)}{(1-\eta)(\gamma+1)}}d\eta =
1751: \frac{\pi}{4}(1-\gamma^2+\gamma \sqrt{\gamma^2-1} ),
1752: \end{equation}
1753: and the integral in the negative arctangent in \eqref{eq:arctrad} is
1754: obviously given by \eqref{eq:eqx5} as well. Using \eqref{eq:arctrad}
1755: and \eqref{eq:eqx5}, we see that the $\alpha$-condition
1756: \eqref{eq:alphabetacond} is equivalent to
1757: $$ \beta^2(\gamma^2-\gamma \sqrt{\gamma^2-1}) = \alpha \iff
1758: 1-2\alpha = \sqrt{1-2\beta^2}, $$
1759: the latter being possible only if $\alpha < 1/2$. In that case
1760: \begin{equation}\label{eq:betaalpha}
1761: \beta = \sqrt{2\alpha(1-\alpha)}.
1762: \end{equation}
1763: Consequently, see \eqref{eq:lambdac},
1764: \begin{equation}\label{eq:lambdavalue}
1765: \lambda = \log \frac{1-\alpha}{\alpha},
1766: \end{equation}
1767: and, see \eqref{eq:hprime},
1768: \begin{equation}\label{eq:hprime2}
1769: h'(s) =
1770: \frac{2}{\pi}\arct\left(\frac{(1-2\alpha)s}{\sqrt{2\alpha(1-\alpha)-s^2}}
1771: \right), \qquad s\in
1772: (-\sqrt{2\alpha(1-\alpha)},\sqrt{2\alpha(1-\alpha)}).
1773: \end{equation}
1774: Furthermore, denoting the integral in \eqref{eq:eqx4} by
1775: $J(\xi,\gamma)$, we easily get
1776: \begin{multline}\label{eq:hnotprime}
1777: \qquad
1778: h(s)= \beta + \int_{-\beta}^s h'(t)dt = \beta(1+J(\xi,\gamma) +
1779: J(-\xi,\gamma) - J(1,\gamma)) \\ =
1780: \frac{2}{\pi}
1781: s\arct\left(\frac{(1-2\alpha)s}{\sqrt{2\alpha(1-\alpha)-s^2}}
1782: \right) +
1783: \frac{\sqrt{2}}{\pi}
1784: \arct
1785: \left(\frac{\sqrt{2(2\alpha(1-\alpha)-s^2)}}{1-2\alpha}\right).
1786: \end{multline}
1787: We have derived a formula for a candidate minimizer, which we now
1788: recognize as the function $\tilde{g}_\alpha$ that we defined in
1789: section 2. To be sure, this function was determined so as to meet the
1790: ramifications of \emph{some} of the constraints. However, looking at
1791: \eqref{eq:hprime2}, we see that $-1 < h'(s) < 1$ for $s\in
1792: (-\beta,\beta)$, so $h$ is indeed $1$-Lipschitz, even though so far we
1793: haven't paid attention to this constraint! Furthermore, since
1794: $h^\prime(s)$ is odd, the constraint \eqref{eq:integral0} is met
1795: automatically, and it is the reason why we were able to satisfy the
1796: boundary constraints $h(-\sqrt{2}/2) = h(\sqrt{2}/2) =
1797: \sqrt{2}/2$. Also, we determined $\beta$ from the requirement that $h$
1798: should satisfy \eqref{eq:alphabetacond}, which under these boundary
1799: conditions is equivalent to the $\alpha$-condition. We conclude that,
1800: at the very least, $\tilde{g}_{\alpha}$ meets all the constraints,
1801: thus is $\alpha$-admissible.
1802:
1803: By Lemma 7, to prove that $\tilde{g}_\alpha$ is the minimizer, it only
1804: remains to prove that $\tilde{g}_{\alpha}$ satisfies the conditions
1805: \eqref{eq:necessarycond}. By \eqref{eq:diffmultline},
1806: $w^\prime(s)\equiv 0$ for $|s|<\beta$. And $w(0) =0$ as $h^\prime(t)$
1807: is odd. So $w(s)\equiv 0$ for $|s|<\beta$, hence the first condition
1808: in \eqref{eq:necessarycond} is met. As for the remaining conditions,
1809: by (anti)symmetry, it suffices to check, say, the third condition,
1810: namely that
1811: $$ F(s,\alpha) := -\int_{-\sqrt{2}/2}^{\sqrt{2}/2}
1812: \tilde{g}_\alpha'(t)
1813: \log|s-t| dt - \lambda(\alpha) s \le 0,\qquad \beta(\alpha) \le s \le
1814: \sqrt{2}/2.
1815: $$
1816: Fix $0<s\le \sqrt{2}/2$, and let $\hat{\alpha}=(1-\sqrt{1-2s^2})/2$,
1817: so that $\beta(\hat{\alpha})=s$. Clearly, because of the first
1818: condition in \eqref{eq:necessarycond}, $F(s,\hat{\alpha})=0$. To
1819: finish the proof, we will now show that $\partial F(s,\alpha)/\partial
1820: \alpha > 0$ for $0<\alpha<\hat{\alpha}$. By \eqref{eq:multline},
1821: \begin{equation}\label{eq:diffbeta}
1822: \frac{\partial F(s,\alpha)}{\partial \beta} = - \int_{-\beta}^\beta
1823: \frac{\partial \tilde{g}_\alpha'(t,\alpha)}{\partial \beta} \log |s-t|
1824: dt - s \frac{\partial \lambda}{\partial \beta}.
1825: \end{equation}
1826: Using \eqref{eq:hprime} and simplifying gives
1827: \begin{equation*}
1828: \frac{\partial \tilde{g}_\alpha'(t)}{\partial \beta} = -\frac{2}{\pi
1829: \beta(1-2 \beta^2)^{1/2}}\cdot \frac{t}{(\beta^2-t^2)^{1/2}}.
1830: \end{equation*}
1831: Since $\beta'(\alpha) = (1-2\beta^2)^{1/2}/\beta$, \eqref{eq:diffbeta}
1832: becomes
1833: \begin{equation*}
1834: \frac{\partial F(s,\alpha)}{\partial \alpha} = \frac{2}{\pi \beta^2}
1835: \int_{-\beta}^\beta \frac{t \log|s-t|}{(\beta^2-t^2)^{1/2}} dt +
1836: \frac{s}{(1-\alpha)\alpha}.
1837: \end{equation*}
1838: Here the integral equals
1839: $$
1840: -(\beta^2-t^2)^{1/2} - \log|s-t|\bigg|_{-\beta}^\beta -
1841: \int_{-\beta}^\beta \frac{(\beta^2-t^2)^{1/2}}{s-t}dt = -\pi
1842: (s-(s^2-\beta^2)^{1/2}),
1843: $$
1844: see \eqref{eq:evaluation}. Therefore
1845: \begin{eqnarray*}
1846: \frac{\partial F(s,\alpha)}{\partial \alpha} &=& -\frac{2}{\beta^2}
1847: \left(s-(s^2-\beta^2)^{1/2}\right) + \frac{s}{(1-\alpha)\alpha} \\
1848: &=& s\left( \frac{1}{(1-\alpha)\alpha}-\frac{2}{\beta^2}\right) +
1849: \frac{2}{\beta^2} (s^2-\beta^2)^{1/2} \\ & = &
1850: \frac{2}{\beta^2} (s^2-\beta^2)^{1/2} > 0.
1851: \end{eqnarray*}
1852:
1853: \subsection{Direct computation of $K(\tilde{g}_\alpha)$}
1854:
1855: Our next goal in this section is to show that
1856: $K(\tilde{g}_\alpha)=-H(\alpha)+\log 2$. There are two ways to do
1857: this. First, looking at the proof of Theorem 8, we see that we may
1858: repeat the arguments of that proof (without assuming the value of
1859: $K(\tilde{g}_\alpha)$ as in that proof) to deduce that the value
1860: $M_\alpha$ of $K(\tilde{g}_\alpha)+H(\alpha)-\log 2$ \emph{must} be
1861: $0$. For, if it were greater than $0$, then, denoting $k=\lfloor
1862: \alpha n^2\rfloor$, we would have
1863: \begin{eqnarray*}
1864: 1 = \pr_n\left( T \in {\cal T}_n \right) &=&
1865: \sum_{\lambda_0\textrm{ of area }k}
1866: \pr_n\bigg( T\in{\cal T}_n : \lambda_T^k =
1867: \lambda_0 \bigg) \\ &\le & p(n^2) \exp\left( - (1+o(1)) n^2 M_{k/n^2}
1868: \right) \xrightarrow[n\to\infty]{} 0
1869: \end{eqnarray*}
1870: (since $M_\alpha$ is obviously continuous in $\alpha$.) On the other
1871: hand, if $M_\alpha < 0$, then for some sufficiently large $n$, we
1872: would have for some diagram $\lambda_0$ of area $\lfloor \alpha n^2
1873: \rfloor$ contained in $\square_n$, that
1874: $K(g_{\lambda_0})+H(\alpha)-\log 2 < 0$ (take a diagram for which
1875: $g_{\lambda_0}$ approximates $\tilde{g}_\alpha$, and use Lemma 2). But
1876: this again implies a contradiction:
1877: $$ 1 \ge \pr_n\left( T \in {\cal T}_n : \lambda_T^{\lfloor \alpha n^2
1878: \rfloor} = \lambda_0 \right) = \exp\left(-(1+o(1))n^2
1879: (K(g_{\lambda_0})+H(\alpha)-\log 2)\right) > 1. $$
1880:
1881: These last remarks notwithstanding, we find it worthwhile to compute
1882: $K(\tilde{g}_\alpha)$ directly, if only to thoroughly test our
1883: derivation of $\tilde{g}_\alpha$, and to show that all the integrals
1884: involved can be evaluated explicitly.
1885:
1886: For $h=\tilde{g}_\alpha$, rewrite \eqref{eq:alphacondition} as
1887: $$ -\int_{-\sqrt{2}/2}^{\sqrt{2}/2} u(h'(u)-\sgn u)du = \alpha. $$
1888: Using this, multiply both sides of \eqref{eq:necessarycond}
1889: by $(h'(s)-\sgn s)$ and integrate, obtaining
1890: \begin{multline}\label{eq:multline5}
1891: K(h) = -\frac{\lambda \alpha}{2} -
1892: \frac{1}{2}\int_{-\sqrt{2}/2}^{\sqrt{2}/2} h'(t) \bigg[ 2t \log|t| -
1893: (t+\sqrt{2}/2)\log |t+\sqrt{2}/2| \\
1894: - (t-\sqrt{2}/2)\log | t-\sqrt{2}/2 | \bigg] dt,
1895: \qquad\qquad\ \ \,
1896: \end{multline}
1897: where we found before that $\lambda = \log((1-\alpha)/\alpha)$.
1898: Denote
1899: $$ S(t) = 2t \log|t| - (t+\sqrt{2}/2)\log |t+\sqrt{2}/2| \\ -
1900: (t-\sqrt{2}/2)\log | t-\sqrt{2}/2 |, $$
1901: and set
1902: $$K_1(h) = \int_{-\sqrt{2}/2}^{\sqrt{2}/2} h'(t)S(t)dt,$$
1903: so that $K(h) = -\lambda\alpha/2 - K_1(h)/2$. Just like
1904: \eqref{eq:diffbeta},
1905: \begin{multline}\label{eq:diffbeta2}
1906: \frac{\partial K_1(h_\alpha)}{\partial \beta} =
1907: \int_{-\sqrt{2}/2}^{\sqrt{2}/2} \frac{\partial h_\alpha'(t)}{\partial
1908: \beta} S(t) dt = \frac{2}{\pi\beta(1-2 \beta^2)^{1/2}}
1909: \int_{-\beta}^\beta \frac{-t}{(\beta^2-t^2)^{1/2}}S(t)dt \\
1910: = -\frac{2}{\pi\beta(1-2\beta^2)^{1/2}} \int_{-\beta}^\beta
1911: (\beta^2-t^2)^{1/2}\bigg[ 2 \log|t|
1912: -\log | t+ \sqrt{2}/2| - \log|t-\sqrt{2}/2| \bigg] dt.
1913: \end{multline}
1914: Denote
1915: $$ E(s,\beta) = \int_{-\beta}^\beta (\beta^2-t^2)^{1/2}\log|t-s| dt,
1916: $$
1917: so that
1918: \begin{equation}\label{eq:diffbeta3}
1919: \frac{\partial K_1(h_\alpha)}{\partial \beta} =
1920: 2E(0,\beta)-E(-\sqrt{2}/2,\beta) - E(\sqrt{2}/2,\beta).
1921: \end{equation}
1922: By \eqref{eq:lambdaterm} and \eqref{eq:evaluation},
1923: \begin{multline}\label{eq:diffs}
1924: \frac{\partial E(s,\beta)}{\partial s} = \int_{-\beta}^\beta
1925: \frac{(\beta^2-t^2)^{1/2}}{s-t}dt \\
1926: = \left\{ \begin{array}{ll} \pi s & |s|< \beta, \\
1927: \pi (\sgn s)\left(|s|-(s^2-\beta^2)^{1/2}\right), &
1928: \beta < |s| < \sqrt{2}/2. \end{array}\right.
1929: \qquad\qquad\qquad\quad\ \
1930: \end{multline}
1931: Therefore
1932: \begin{equation}\label{eq:EEE}
1933: 2 E(0,\beta) = E(\beta,\beta)+E(-\beta,\beta)+ \pi\int_\beta^0 s ds +
1934: \pi \int_{-\beta}^0 s ds = E(\beta,\beta)+ E(-\beta,\beta)-\pi
1935: \beta^2.
1936: \end{equation}
1937: Likewise
1938: \begin{equation}\label{eq:EEE2}
1939: E(-\sqrt{2}/2,\beta)+E(\sqrt{2}/2,\beta) =
1940: E(-\beta,\beta)+E(\beta,\beta) + 2\pi \int_\beta^{\sqrt{2}/2} \left(
1941: s-(s^2-\beta^2)^{1/2}\right) ds,
1942: \end{equation}
1943: where
1944: \begin{multline}\label{eq:multline6}
1945: \int_\beta^{\sqrt{2}/2} (s^2-\beta^2)^{1/2}ds = \frac{1}{2}\left[
1946: s(s^2-\beta^2)^{1/2}-\beta^2 \log\left( s+(s^2-\beta^2)^{1/2} \right)
1947: \right]\bigg|_\beta^{\sqrt{2}/2} \\
1948: = \frac{1}{2}\left( \frac{1-2\alpha}{2}-\alpha(1-\alpha)\log
1949: \frac{1-\alpha}{\alpha} \right).
1950: \end{multline}
1951: So, using $\beta=(2\alpha(1-\alpha))^{1/2}$,
1952: $$
1953: E(-\sqrt{2}/2,\beta)+E(\sqrt{2}/2,\beta)=E(-\beta,\beta)+E(\beta,\beta)
1954: + \pi\left(-\beta^2+\alpha+\alpha(1-\alpha)\log\frac{1-\alpha}{\alpha}
1955: \right),
1956: $$
1957: and, combining this relation with \eqref{eq:EEE}, we simpify
1958: \eqref{eq:diffbeta3} to
1959: $$
1960: \frac{\partial K_1(h_\alpha)}{\partial \beta} = -\pi\left(
1961: \alpha+\alpha(1-\alpha) \log\frac{1-\alpha}{\alpha} \right).
1962: $$
1963: So, by \eqref{eq:diffbeta2}
1964: $$
1965: \frac{\partial K_1(h_\alpha)}{\partial \alpha} =
1966: \frac{\partial K_1(h_\alpha)}{\partial \beta}\cdot
1967: \frac{(1-2\beta^2)^{1/2}}{\beta} = \frac{1}{1-\alpha} +
1968: \log\frac{1-\alpha}{\alpha}. $$
1969: Since $h_\alpha' \equiv 0$ at $\alpha=1/2$, we have $K_1(h)=0$ at
1970: $\alpha=1/2$. Hence
1971: \begin{multline}
1972: K_1(h_\alpha) = \int_{1/2}^\alpha \left(\frac{1}{1-x}+\log
1973: \frac{1-x}{x} \right) dx = -\log(1-\alpha)-2\log
1974: 2 \\ -(1-\alpha)\log(1-\alpha) -\alpha\log\alpha, \qquad\qquad\qquad
1975: \end{multline}
1976: which gives finally for $K(h_\alpha)$
1977: $$ K(h) = \alpha\log \alpha+(1-\alpha)\log(1-\alpha)+\log 2 =
1978: -H(\alpha) + \log 2.
1979: $$
1980: The proof of Theorem 7 is complete.
1981: \qed
1982:
1983: \subsection{The parametric family $\tilde{g}_\alpha$}
1984:
1985: The minimality proof in section 3.2 relied on the possibility to
1986: consider simultaneously the whole family of variational problems, and
1987: thus to differentiate the minimizer $\tilde{g}_{\alpha}$ with respect
1988: to $\alpha$. Moreover, to reveal a little secret, we anticipated the
1989: formula \eqref{eq:lambdavalue} for the Lagrange multiplier
1990: $\lambda$. According to a general (semiformal) recipe of the calculus
1991: of variations (more specifically, mathematical programming), we
1992: knew that this $\lambda$, dual to the $\alpha$-condition, should be
1993: equal to $d K(\tilde{g}_\alpha)/d\alpha$, which we have proved to
1994: be correct. The advantages of this approach of varying the parameter
1995: $\alpha$ go even deeper than that. It will turn out that the partial
1996: derivative of the minimizer $g_{\alpha}(\cdot)$ with respect to
1997: $\alpha$ is the key to the distributional properties of the random
1998: tableau. Using the formula for the minimizer, we compute easily that
1999: \begin{equation}\label{eq:partialdiff}
2000: \frac{\partial \tilde{g}_\alpha(u)}{\partial \alpha} = \left\{
2001: \begin{array}{lll}
2002: 0 & \qquad\qquad\quad\
2003: & \sqrt{2\alpha(1-\alpha)} < |u| \le \sqrt{2}/2 \\
2004: \frac{\sqrt{2\alpha(1-\alpha)-u^2}}{\pi\alpha(1-\alpha)} & &
2005: |u| \le \sqrt{2\alpha(1-\alpha)} \end{array} \right.
2006: \end{equation}
2007: For each $\alpha$, direct integration reveals that
2008: $\partial\tilde{g}_\alpha(u)/\partial\alpha$ is a probability density
2009: function, i.e.
2010: $$ \int_{-\sqrt{2}/2}^{\sqrt{2}/2}
2011: \frac{\partial\tilde{g}_\alpha(u)}{\partial \alpha}\,du = 1. $$
2012: (In fact, it is the density of the semicircle distribution, and it will play
2013: a prominent role later. See sections 5, 8.1, 8.2.) This observation is in
2014: perfect harmony with the fact that $\tilde{g}_{\alpha}$ satisfies the
2015: $\alpha$-condition, thus providing a partial check of our computations. Indeed
2016: \begin{eqnarray*}
2017: \int_{-\sqrt{2}/2}^{\sqrt{2}/2} \left(\tilde{g}_\alpha(u)-|u|\right)du
2018: &=& \int_{-\sqrt{2}/2}^{\sqrt{2}/2}
2019: \left(\tilde{g}_\alpha(u)-\tilde{g}_0(u)\right)du
2020: \qquad\qquad\qquad\qquad\qquad\qquad\qquad
2021: \\ & \hspace{-120.0pt}=& \hspace{-60.0pt}
2022: \int_{-\sqrt{2}/2}^{\sqrt{2}/2} \left(\int_0^\alpha
2023: \frac{\partial\tilde{g}_s(u)}{\partial s}ds\right)du =
2024: \int_0^\alpha \left( \int_{-\sqrt{2}/2}^{\sqrt{2}/2}
2025: \frac{\partial\tilde{g}_s(u)}{\partial s} du\right) ds \\
2026: & \hspace{-120.0pt}=& \hspace{-60.0pt}
2027: \int_0^\alpha 1\, ds = \alpha.
2028: \end{eqnarray*}
2029: Had we been presented with the minimizer $\tilde{g}_{\alpha}$ ``out of
2030: the blue'', this would have been the least computational way to prove
2031: its $\alpha$-admissibility.
2032:
2033: \section{The boundary of the square}
2034:
2035: \subsection{Proof of Theorem 3}
2036:
2037: In this section, we prove Theorem 3. As was remarked in section 1.3,
2038: the RSK correspondence induces a correspondence between minimal
2039: Erd\"os-Szekeres permutations $\pi$ of $1,2,\ldots,n^2$ and pairs
2040: $T_1,T_2\in {\cal T}_n$ of square tableaux. By the well known result
2041: of Schensted \cite{schensted}, in this correspondence the length
2042: $l_{n,k}$ of the longest increasing subsequence in
2043: $\pi(1),\pi(2),\ldots,\pi(k)$ is equal to the
2044: length
$\lambda_{T_1}^k(1)$ of the first row of $\lambda_{T_1}^k$. So
2045: the
2046: distribution of $l_{n,k}$ under a uniform random choice of
2047: minimal
Erd\"os-Szekeres permutation $\pi$ is equal to the
2048: distribution of the
length of the first row of $\lambda_T^k$ in a
2049: uniform random square
tableau $T\in{\cal T}_n$. Denoting for the
2050: remainder of this section
$\alpha=\alpha(k)=k/n^2$, we can therefore
2051: reformulate Theorem 3 as
stating that
2052: \begin{equation}\label{eq:boundary}
\max_{\alpha_0\le k/n^2\le
2053: 1/2}
\mathbb{P}_n\left( T\in{\cal T}_n :
\left|\lambda_T^{k}(1) -
2054: 2\sqrt{\alpha(1-\alpha)}n\right| > \alpha_0^{1/2}\omega(n)n \right)
2055: \xrightarrow[n\to\infty]{} 0.
\end{equation}
Theorem 8 looks as if
2056: it might imply \eqref{eq:boundary}. In fact, it
only implies a lower
2057: bound on $\lambda_T^k(1)$. The reason is that
$g_{\lambda_T^k}$ can
2058: be very close in the supremum norm to
$\tilde{g}_\alpha$ (as is known
2059: to happen with high probability by
Theorem 8), while $n^{-1}
2060: \lambda_T^k(1)$ might still be much larger
than $2
2061: \sqrt{\alpha(1-\alpha)}$ (see \eqref{eq:firstrow}
2062: below).
2063:
2064: \paragraph{Lemma 9.} Let $\alpha_0=n^{-2/3+\epsilon}$,
2065: $\delta=n^{-1/3(1-\epsilon)
}$, $\epsilon\in (0,2/3)$. Then
2066:
\begin{equation}\label{eq:boundaryweak}
\mathbb{P}_n\left(
2067: T\in{\cal T}_n :
\min_{\alpha_0\le\alpha\le 1/2} (\lambda_T^{k}(1)
2068: -2\sqrt{\alpha(1-\alpha)}n)
\le -\delta n \right) = O(n^{-b})
2069: \end{equation}
for every $b>0$.
\paragraph{Proof.} We use the
2070: notation of Theorem 8. The length of the
first row $\lambda_T^k(1)$
2071: can be extracted from the rotated
coordinate graph $g_{\lambda_T^k}$
2072: using the following
2073: relation:
\begin{equation}\label{eq:firstrow}
\frac{1}{n}\lambda_T^k(1)
2074: = \sqrt{2}\, \inf\left\{ u \in [0,\sqrt{2}/2]
: g_{\lambda_T^k}(u) =
2075: u \right\}.
\end{equation}
It follows from \eqref{eq:hprime2} that,
2076: uniformly for $\alpha\in [\alpha_0,1/2]$
and $|u|<
2077: \sqrt{2\alpha(1-\alpha)}$,
$$
|\partial\tilde{g}_{\alpha}(u)/\partial
2078: u
2079: -1|=\frac{2}{\pi}\tan^{-1}\frac{\sqrt
{2\alpha(1-\alpha)-u^2}}{(1-2\alpha)|u|}\ge
2080: c(\sqrt{2\alpha(1-\alpha)}-|u|),
$$
$c>0$ being an absolute
2081: constant. Consequently, for $\alpha\in
2082: [\alpha_0,1/2]$,
$$
\tilde{g}_\alpha(\sqrt{2\alpha(1-\alpha)}-\delta)-(
\sqrt{2\alpha(1-\alpha)}-\delta)
2083: \ge c\delta^{3/2}.$$
So if $T\in {\cal T}_n$ has the property that,
2084: for some $k$ in question,
$$
\lambda_T^{k} (1) -
2085: 2\sqrt{\alpha(1-\alpha)}n
< -\delta n, $$
then by
2086: \eqref{eq:firstrow},
\begin{eqnarray*}
||g_{\lambda_T^{k
}}-\tilde{g}_\alpha||_\infty
2087: &\ge& \sup \{
|g_{\lambda_T^{k}}(u)-\tilde{g}_\alpha(u)|
2088: :
\sqrt{2\alpha(1-\alpha)}-\delta < u < \sqrt{2\alpha(1-\alpha)} \}
2089: \\
& = & \sup \{ \tilde{g}_\alpha(u)-u
2090: :
\sqrt{2\alpha(1-\alpha)}-\delta< u < \sqrt{2\alpha(1-\alpha)} \}
2091: \ge
c\delta^{3/2}.
\end{eqnarray*}
So, by Theorem 8 with
2092: $\epsilon:=c\delta^{3/2}$,
2093: \begin{multline*}
2094: \qquad\qquad \mathbb{P}_n\left( T\in{\cal T}_n :
2095: \min_{\alpha_0\le\alpha\le 1/2}(\lambda_T^{k} (1) -
2096: 2\sqrt{\alpha(1-\alpha)}n)
2097: \le -\delta n \right) \\
2098: \le \mathbb{P}_n\left( T\in{\cal T}_n : \max_{\alpha_0\le\alpha\le
2099: 1/2} || g_{\lambda_T^{k}}-\tilde{g}_\alpha||_\infty \ge
2100: c\delta^{3/2}\right) \\ \le \exp(3n-\hat{c}n^2\delta^{3r/2})
2101: \le \exp(3n-\hat{c}n^{2-r(1-\epsilon)/2}) \xrightarrow[n\to\infty]{}
2102: 0,\qquad\qquad\
2103: \end{multline*}
2104: provided that we choose a feasible $r$, i. e. $r\in (2,3)$, such
2105: that $r<2(1-\epsilon)^{-1}$.
2106: \qed
2107:
2108: \bigskip
2109: To prove the upper bound and thus conclude the proof of Theorem 3, it
suffices to prove an upper bound for the \emph{expected value} of
$\lambda_T^k$, namely that, for $\alpha_0\le\alpha\le 1/2$,
\begin{equation}\label{eq:expected}
\mathbb{E}_n \left[\lambda_T^{k}(1) \right] \le 2\sqrt{\alpha(1-\alpha)}n +
O(\alpha_0^{1/2}n),
\end{equation}
where $\mathbb{E}_n$ denotes expectation with respect to the
probability measure $\mathbb{P}_n$.
Indeed, choosing $\omega(n)\to\infty$ however slowly, we bound
$$
\mathbb{P}_n\left(T\in{\cal T}_n : \lambda_T^{k} (1)
\ge 2\sqrt{\alpha(1-\alpha)}n+ \alpha_0^{1/2}\omega(n)n\right)
\qquad\qquad\qquad\qquad\qquad\qquad\qquad $$
\vspace{-27.0pt}
\begin{eqnarray*}
\textrm{(by Markov's inequality)}
& \le & (\alpha_0^{1/2}\omega(n)n)^{-1}\mathbb{E}_n
\left[\max(0,\lambda_T^k(1)-2\sqrt{\alpha(1-\alpha)}n)\right]\\
\textrm{(by Lemma 9, for any $b>0$)}
& \le & (\alpha_0^{1/2}\omega(n)n)^{-1}
\left(\mathbb{E}_n\left[\lambda_T^k(1)-
2\sqrt{\alpha(1-\alpha)}+\delta n\right]+O(n^{1-b}) \right)\\
& = & O((\alpha_0^{1/2}n+\delta n)/(\alpha_0^{1/2}\omega(n)n))=O(\omega(n)^{-1}).
\end{eqnarray*}
\bigskip
Write
$$ \lambda_T^k(1) = \sum_{j=1}^k I_{n,j}, $$
where
$I_{n,j} = \lambda_T^j(1)-\lambda_T^{j-1}(1) = $
indicator of the event that $\lambda_T^j$ is obtained from
$\lambda_T^{j-1}$ by adding a box to the first row. Let $p_{n,j} =
\mathbb{E}_n(I_{n,j})$.
\paragraph{Lemma 10.} In the notation of Lemma 9, as $n\to\infty$,
$$ p_{n,j} \le \frac{n^2-2j}{n\sqrt{j(n^2-j)}}+
O(\delta n(n^2-2j+1)^{-1}), $$
uniformly for $\alpha_0\le j/n^2\le 1/2$.
\paragraph{Proof.} Let ${\cal Y}_{n,j}$ be the set of Young diagrams
2110: of area $j$ contained in the $n\times n$ square. For a diagram
2111: $\lambda \in {\cal Y}_{n,j}$, denote by $\textrm{next}(\lambda)$ the
2112: diagram obtained from $\lambda$ by adding a box to the first
2113: row. Then, conditioning $I_{n,j}$ on the shape $\lambda_T^{j-1}$, we
2114: write
2115: \begin{eqnarray*}
2116: p_{n,j} &=& \mathbb{P}_n\left( \lambda_T^j =
2117: \textrm{next}(\lambda_T^{j-1})\right) = \sum_{ \lambda\in {\cal Y}_{n,j-1}}
2118: \frac{d(\lambda) d(\square_n\setminus
2119: \textrm{next}(\lambda))}{d(\square_n)} \\ &=&
2120: \sum_{\lambda \in {\cal Y}_{n,j-1}}
2121: \frac{d(\textrm{next}(\lambda))d(\square_n\setminus
2122: \textrm{next}(\lambda))}{d(\square_n)}\cdot
2123: \frac{d(\lambda)}{d(\textrm{next}(\lambda))}
2124: \end{eqnarray*}
2125: This is nearly an average over ${\cal Y}_{n,j}$ with respect to the
2126: measure \eqref{eq:dmeasure}; in fact, slightly less, since not any
2127: $\lambda'\in {\cal Y}_{n,j}$ is of the form $\textrm{next}(\lambda)$
2128: for some $\lambda\in{\cal Y}_{n,j-1}$. It follows from the convexity
2129: of the function $x\to x^2$ that
2130: \begin{multline}\label{eq:beforeamus}
2131: \qquad\qquad
2132: p_{n,j}^2 \le \sum_{\lambda \in {\cal Y}_{n,j-1}}
2133: \frac{d(\textrm{next}(\lambda))d(\square_n\setminus
2134: \textrm{next}(\lambda))}{d(\square_n)}\cdot \left(
2135: \frac{d(\lambda)}{d(\textrm{next}(\lambda))} \right)^2 \\ =
2136: \sum_{\lambda \in {\cal Y}_{n,j-1}}
2137: \frac{d(\lambda)d(\square_n\setminus \lambda)}{d(\square_n)}\cdot
2138: \frac{d(\lambda)d(\square_n\setminus \textrm{next}(\lambda))}
2139: {d(\textrm{next}(\lambda))d(\square_n\setminus \lambda)}.
2140: \qquad\qquad\qquad\qquad\
2141: \end{multline}
2142: We now note the amusing identity
2143: \begin{equation}\label{eq:amus}
2144: \frac{d(\lambda)d(\square_n\setminus \textrm{next}(\lambda))}
2145: {d(\textrm{next}(\lambda))d(\square_n\setminus \lambda)}
2146: = \frac{n^2-\lambda(1)^2}{j(n^2-j+1)}, \qquad (\lambda \in {\cal Y}_{n,j-1})
2147: \end{equation}
2148: which follows from writing out the hook products for $d(\cdot)$ in
2149: \eqref{eq:hook} and observing cancellation of almost all the factors -
2150: see Figure 5. Here is a proof of \eqref{eq:amus}. Clearly the only
2151: hook lengths influenced by this operation are of the cells in the
2152: first row and the $(\lambda(1)+1)$-th column. In particular,
2153: $$ \frac{d(\lambda)}{d(\textrm{next}(\lambda))} = \frac{1}{j}
2154: \prod_{i=1}^{\lambda(1)}
2155: \frac{\lambda(1)-i+1+\lambda'(i)}{\lambda(1)-i+\lambda'(i)};
2156: $$
2157: here $\lambda'(i)$ is the number of cells in the $i$-th column of
2158: $\lambda$. Clearly the fraction factors ``telescope'' on each
2159: subinterval of $[1,\lambda(1)]$ where $\lambda'(i)$ is constant. Let
2160: $[i_1,i_2]$ be such a (maximal) subinterval. Maximality implies that
2161: $(i_2,\lambda'(i_2))$ is a corner of $\lambda$, and that
2162: $(\lambda'(i_1)+1,i_1)$ is a corner of $\square_n\setminus
2163: \textrm{next}(\lambda)$. Then
2164: $$
2165: \prod_{i=1}^{\lambda(1)}
2166: \frac{\lambda(1)-i+1+\lambda'(i)}{\lambda(1)-i+\lambda'(i)} =
2167: \frac{\lambda(1)-i_1+1+\lambda'(i_1)}{\lambda(1)-i_2+\lambda'(i_2)}
2168: =
2169: \frac{h_{\square_n\setminus\textrm{next}(\lambda)}(\lambda'(i_1)+1,\lambda(1)+1)}{h_\lambda(1,u_2)}
2170: $$
2171: where, say, $h_\lambda(u,v)$ denotes the hook length for a cell $(u,v)
2172: \in \lambda$. Multiplying these fractions for all such subintervals
2173: $[i_1,i_2]$, we get
2174: \begin{equation} \label{eq:amus2}
2175: \frac{d(\lambda)}{d(\textrm{next}(\lambda))} = \frac{1}{j}
2176: \left(\prod_{(u,v)\in \textrm{corners}(\lambda)} f(u,v)\right)^{-1} \cdot
2177: \left(\prod_{(u,v)\in \textrm{corners}(\square_n\setminus \lambda)}
2178: g(u,v) \right).
2179: \end{equation}
2180: Here $\textrm{corners}(\mu)$ is the corner set of a diagram $\mu$;
2181: $f(u,v)$ is the hook length of a cell in the first row of $\lambda$
2182: whose vertical leg ends at the corner $(u,v)\in
2183: \textrm{corners}(\lambda)$; $g(u,v)$ is the hook length of a cell in
2184: $\square_n\setminus\textrm{next}(\lambda)$ from the
2185: $(\lambda(1)+1)$-th column whose horizontal arm ends at the corner
2186: $(u,v) \in \textrm{corners}(\square_n\setminus \lambda)$. Next,
2187: considering separately the first row cells $(1,v)$, $v>\lambda(1)$,
2188: the top $\lambda'(1)$ cells in the $(\lambda(1)+1)$-th column, and
2189: finally the bottom $n-\lambda'(1)$ cells in that column, we obtain
2190: \begin{equation}\label{eq:amus3}
2191: \frac{d(\square_n\setminus
2192: \textrm{next}(\lambda))}{d(\square_n\setminus \lambda)} =
2193: \frac{n-\lambda(1)}{n^2-j+1} \cdot \prod_{k=2}^{\lambda'(1)}
2194: \frac{\lambda(1)-\lambda(k)+k}{\lambda(1)-\lambda(k)+k-1} \cdot
2195: \frac{\lambda(1)+n}{\lambda(1)+\lambda'(1)}.
2196: \end{equation}
2197: Here, analogously to the $d(\lambda)/d(\textrm{next}(\lambda))$ case,
2198: \begin{multline}\label{eq:amus4}
2199: \qquad
2200: \frac{1}{\lambda(1)+\lambda'(1)} \prod_{k=2}^{\lambda'(1)}
2201: \frac{\lambda(1)-\lambda(k)+k}{\lambda(1)-\lambda(k)+k-1} \\ =
2202: \left(\prod_{(u,v)\in \textrm{corners}(\lambda)} f(u,v)\right) \cdot
2203: \left(\prod_{(u,v)\in \textrm{corners}(\square_n\setminus \lambda)}
2204: g(u,v) \right)^{-1}. \qquad
2205: \end{multline}
2206: Putting \eqref{eq:amus2}, \eqref{eq:amus3}, \eqref{eq:amus4} together
2207: gives \eqref{eq:amus}.
2208:
2209: \begin{figure}[h!]
2210: \begin{center}
2211: \begin{picture}(200,200)(0,0)
2212: \setlength{\unitlength}{1.7 pt}
2213: \multiput(0,100)(0,-10){10}{\framebox(10,10)}
2214: \multiput(10,100)(0,-10){10}{\framebox(10,10)}
2215: \multiput(20,100)(0,-10){9}{\framebox(10,10)}
2216: \multiput(30,100)(0,-10){7}{\framebox(10,10)}
2217: \multiput(40,100)(0,-10){7}{\framebox(10,10)}
2218: \multiput(50,100)(0,-10){4}{\framebox(10,10)}
2219: \multiput(60,100)(0,-10){1}{\framebox(10,10)}
2220: \multiput(62,101)(2,0){4}{\line(0,1){8}}
2221: \put(0,10){\framebox(100,100)}
2222: \multiput(60,90)(0,-10){9}{\dashbox{2}(10,10)}
2223: \multiput(70,100)(10,0){3}{\dashbox{2}(10,10)}
2224: \multiput(1,101)(10,0){6}{\line(1,1){8}}
2225: \multiput(71,101)(10,0){3}{\line(1,1){8}}
2226: \multiput(61,91)(0,-10){9}{\line(1,1){8}}
2227: \put(0,106){\scriptsize{15}}
2228: \put(5,101){\scriptsize{16}}
2229: \put(11,106){\scriptsize{14}}
2230: \put(15,101){\scriptsize{15}}
2231: \put(21,106){\scriptsize{12}}
2232: \put(25,101){\scriptsize{13}}
2233: \put(33,106){\scriptsize{9}}
2234: \put(35,101){\scriptsize{10}}
2235: \put(43,106){\scriptsize{8}}
2236: \put(46,101){\scriptsize{9}}
2237: \put(53,106){\scriptsize{4}}
2238: \put(56,101){\scriptsize{5}}
2239: \put(73,106){\scriptsize{2}}
2240: \put(76,101){\scriptsize{1}}
2241: \put(83,106){\scriptsize{3}}
2242: \put(86,101){\scriptsize{2}}
2243: \put(93,106){\scriptsize{4}}
2244: \put(96,101){\scriptsize{3}}
2245: \put(63,96){\scriptsize{2}}
2246: \put(66,91){\scriptsize{1}}
2247: \put(63,86){\scriptsize{3}}
2248: \put(66,81){\scriptsize{2}}
2249: \put(63,76){\scriptsize{4}}
2250: \put(66,71){\scriptsize{3}}
2251: \put(63,66){\scriptsize{6}}
2252: \put(66,61){\scriptsize{5}}
2253: \put(63,56){\scriptsize{7}}
2254: \put(66,51){\scriptsize{6}}
2255: \put(63,46){\scriptsize{8}}
2256: \put(66,41){\scriptsize{7}}
2257: \put(61,36){\scriptsize{11}}
2258: \put(65,31){\scriptsize{10}}
2259: \put(61,26){\scriptsize{12}}
2260: \put(65,21){\scriptsize{11}}
2261: \put(61,16){\scriptsize{14}}
2262: \put(65,11){\scriptsize{13}}
2263: \thicklines
2264: \put(0,10){\line(1,0){20}}
2265: \put(20,10){\line(0,1){10}}
2266: \put(20,20){\line(1,0){10}}
\put(30,20){\line(0,1){20}}
\put(30,40){\line(1,0){20}}
\put(50,40){\line(0,1){30}}
\put(50,70){\line(1,0){10}}
\put(60,70){\line(0,1){40}}
\end{picture}
\caption{Illustration of \eqref{eq:amus} for
$\lambda=(6,6,6,6,5,5,5,3,3,2)$: The numbers in the cells are the hook
lengths before and after the new cell is added.}
\end{center}
\end{figure}
Combining \eqref{eq:beforeamus} and \eqref{eq:amus} gives that
\begin{equation}\label{eq:boundpnj2}
p_{n,j}^2 \le \mathbb{E}_n\left[
\frac{n^2-\lambda_T^{j-1}(1)^2}{j(n^2-j+1)}\right]
\end{equation}
By Lemma 9, we may write
$$
\mathbb{E}_n(\lambda_T^{j-1}(1)) \ge
\frac {2\sqrt{j(n^2-j)}}{n}-\delta n,$$
($\delta=n^{-(1-\epsilon)/3}$), for all $j/n^2\in [\alpha_0,1/2]$.
So,
using $\mathbb{E}_n^2[\lambda_T^{j-1}(1)]\le
\mathbb{E}[(\lambda_T^{j-1}(1))^2]$,
$$ p_{n,j}^2 \le \frac{(n^2-2j)^2}{n^2\cdot j(n^2-j)}+\frac{4\delta}{\sqrt{j(n^2-j)}},
$$
or, using $(1+z)^{1/2}\le 1+z/2$ for $j<n^2/2$,
$$ p_{n,j}\le \frac{n^2-2j}{n\sqrt{j(n^2-j)}}+O(\delta n (n^2-2j+1)^{-1}).
$$
The estimate holds for $j=n^2/2$ as well, since
$\delta^{1/2}n^2\to\infty$.
\qed
\bigskip
Note that \eqref{eq:boundpnj2} implies in particular the rough bound
$$
p_{n,j} \le \frac{n}{\sqrt{j (n^2-j+1)}},
$$
valid for all $j\le n^2$. Now, to complete the proof of Theorem 3, we use this bound for
$j \le \alpha_0 n^2$ and Lemma 10 for $j>\alpha_0 n^2$. First
$$
\mathbb{E}_n\bigg[\lambda_T^{k}(1)\bigg] =
\sum_{j\le \alpha_0n^2} p_{n,j} +
\sum_{\alpha_0n^2<j\le k}p_{n,j}=\Sigma_1+\Sigma_2.$$
Here
$$
\Sigma_1\le 2\sum_{j\le\alpha_0n^2}j^{-1/2}=O(n\alpha_0^{1/2}),
$$
and
$$
\Sigma_2\le\sum_{\alpha_0n^2<j\le k}\frac{n^2-2j}{n\sqrt{j(n^2-j)}}+
O(\delta n\log n).
$$
The last sum is bounded above by
$$
n\int_{\alpha_0-n^{-2}}^{\alpha}\frac{1-2t}{\sqrt{t(1-t)}}\,dt=2n\sqrt{\alpha(1-\alpha)}
+O(n\alpha_0^{1/2}).
$$
Therefore, since $\alpha_0^{1/2}\gg \delta\log n$,
$$
\mathbb{E}_n[\lambda_T^k]\le 2n\sqrt{\alpha(1-\alpha)}+O(n\alpha_0^{1/2}).
$$
2267: So \eqref{eq:expected} follows. Theorem 3 is proved.
2268: \qed
2269:
2270: %To complete the proof of Theorem 3, now use Lemma 9, taking $\alpha_0$
2271: %very small ???, to write
2272: %\begin{eqnarray*}
2273: %\mathbb{E}_n\bigg[\lambda_T^{\lfloor \alpha n^2 \rfloor}(1)\bigg] &=&
2274: %\sum_{j=1}^{\lfloor \alpha n^2\rfloor} p_{n,j} \le (1+o(1))
2275: %\sum_{j=1}^{\lfloor \alpha n^2\rfloor} \frac{n^2-2j}{n\sqrt{j(n^2-j)}}
2276: %\\ &=& (1+o(1))
2277: %n \sum_{j=1}^{\lfloor \alpha n^2\rfloor} \frac{
2278: %1-\frac{2j}{n^2}}{\sqrt{\frac{j}{n^2}
2279: %\left(1-\frac{j}{n^2}\right) }}\cdot \frac{1}{n^2} =
2280: %(1+o(1))n \int_0^\alpha \frac{1-2t}{\sqrt{t(1-t)}}dt
2281: %\\ &=& (1+o(1))2\sqrt{\alpha(1-\alpha)} n,
2282: %\end{eqnarray*}
2283: %which is \eqref{eq:expected}. Theorem 3 is proved.
2284: %\qed
2285:
2286: \subsection{Proof of Theorem 1(i)}
2287:
2288: With our enhanced understanding of the distribution of
2289: $\lambda_T^k(1)$, we may now prove Theorem 1(i). First we show that
2290: for individual boundary points, the tableau approaches the limit
2291: surface. Fix $(x,y)$ on the boundary of the square. By symmetry, we
2292: may assume that $y=0, 0<x<1$. Let $\epsilon>0$. Denote
2293: $$ \alpha = L(x,0) = \frac{1-\sqrt{1-x^2}}{2}, $$ so that
2294: $x=2\sqrt{\alpha(1-\alpha)}$. For any tableau $T \in {\cal T}_n$,
2295: denote $k_T = t_{\lfloor nx\rfloor+1,1}$, and let $\beta_T =
2296: k_T/n^2$. We want to show that with high probability,
2297: $|\beta_T-\alpha|$ is small. Note that $k_T$ is an integer
2298: representing the smallest $j$ for which $\lambda_T^j > nx$. Therefore
2299: $nx\le \lambda_T^{k_T}(1) < nx+1$, or
2300: \begin{equation}\label{eq:stam1}
2301: \left| \lambda_T^{k_T}(1)-x \right| \le \frac{1}{n}
2302: \end{equation}
2303: The function $f(t):=L(t,0)=(1-\sqrt{1-t^2})/2$ is monotonically
2304: increasing and uniformly continuous on $[0,1]$. Choose a $\delta>0$
2305: such that $|t-t'|<\delta$ implies $|f(t)-f(t')|<\epsilon/3$. Choose
2306: numbers $0=a_0<a_1<a_2<\ldots<a_N=1/2$ such that $a_{i+1}-a_i <
2307: \epsilon/3$,\ $i=0,1,2,\ldots,N-1$. Denote
2308: $x_i=f^{-1}(a_i)=2\sqrt{a_i(1-a_i)}$.
2309:
2310: Let $T\in {\cal T}_n$ be a tableau that satisfies
2311: \begin{equation}\label{eq:stam2}
2312: \left| \frac{1}{n}\lambda_T^{\lfloor a_i n^2\rfloor}(1) - x_i\right| <
2313: \frac{\delta}{2}, \qquad (i=1,2,\ldots,N)
2314: \end{equation}
2315: (this happens with high probability, by \eqref{eq:boundary}). Let
2316: $0\le i<N$ be such that $a_i \le \beta_T < a_{i+1}$. Then clearly
2317: \begin{equation}\label{eq:stam3}
2318: x_i - \frac{\delta}{2} < \frac{1}{n}\lambda_T^{\lfloor a_i n^2 \rfloor}(1)
2319: \le \frac{1}{n}\lambda_T^{k_T}(1) \le
2320: \frac{1}{n}\lambda_T^{\lfloor a_{i+1} n^2\rfloor}(1) < x_{i+1}+
2321: \frac{\delta}{2}
2322: \end{equation}
2323: Combining this with \eqref{eq:stam1} we get, for $n>2/\delta$,
2324: $$ x_i - \delta < x < x_{i+1} + \delta.$$
2325: Therefore
2326: $$ a_i - \frac{\epsilon}{3} < \alpha=f(x) < a_{i+1}+\frac{\epsilon}{3},
2327: $$
2328: and, since also $a_i \le \beta_T < a_{i+1}$ and $a_{i+1}-a_i <
2329: \epsilon/3$, we get
2330: $$ |\beta_T-\alpha|<\epsilon. $$
2331: Summarizing, we have shown that
2332: \begin{multline}\label{eq:boundary2}
2333: \mathbb{P}_n\left( T \in {\cal T}_n : |\beta_T-\alpha|<\epsilon
2334: \right) \\ \ge
2335: \mathbb{P}_n\left( T\in {\cal T}_n : \forall\ i=1,2,\ldots,N,\ \ \left|
2336: \frac{1}{n}\lambda_T^{\lfloor a_i n^2\rfloor}(1)-x_i\right| <
2337: \frac{\delta}{2} \right)
2338: \xrightarrow[n\to\infty]{} 1.
2339: \end{multline}
2340: Theorem 1(i) now follows easily. It is enough to say that, because of
2341: the monotonicity of the tableau $t_{i,j}$ as a function of $i$ and
2342: $j$, and the monotonicity of the limit surface function $L$,
2343: given $\epsilon>0$ we can find finitely many points
2344: $(x_1,y_1),(x_2,y_2),\ldots,(x_N,y_N) \in [0,1]\times[0,1]\setminus
2345: \{(0,0),(1,0),(0,1),(1,1)\} $ such that the event inclusion
2346: \begin{multline}\label{eq:union}
2347: \qquad
2348: \bigg\{ T\in{\cal T}_n : \max_{1\le i,j\le n}
2349: \left|\frac{1}{n^2}t_{i,j} -
2350: L\left(\frac{i}{n},\frac{j}{n}\right)\right| > \epsilon \bigg
2351: \} \subseteq \\ \bigcup_{i=1}^N
2352: \bigg\{ T \in {\cal T}_n : \left| \frac{1}{n^2}
2353: t_{\lfloor n x_i\rfloor +1, \lfloor n y_i\rfloor+1} - L(x_i,y_i)
2354: \right| > \frac{\epsilon}{10}
2355: \bigg\} \qquad\qquad\qquad
2356: \end{multline}
2357: holds. But now, the $\pr_n$-probability of each of the individual
2358: events in this union tends to 0 as $n\to\infty$ -- because of Theorem
2359: 1(ii) for the points $(x_i,y_i)$ in the interior of the square (using
2360: the continuity of the function $L$), and because of
2361: \eqref{eq:boundary2} for the points on the boundary. \qed
2362:
2363: \section{The hook walk and the cotransition measure of a diagram}
2364:
2365: In this section, we study the location of the $k$-th entry in the
2366: random tableau $T \in {\cal T}_n$, when $k\approx \alpha\cdot
2367: n^2$. The idea is to condition the distribution of the location of the
2368: $k$-th entry on the shape $\lambda_T^k$ of the $k$-th subtableau of
2369: $T$. Given the shape $\lambda_T^k$, the distribution of the location
2370: of the $k$-th entry is exactly the so-called \emph{cotransition
2371: measure} of $\lambda_T^k$ (see below). We know from Theorem 8 that
2372: with high probability, the rescaled shape of $\lambda_T^k$ is
2373: approximately described in rotated coordinates by the level curve
2374: $v=\tilde{g}_\alpha(u)$. Romik \cite{romik} showed that the
2375: cotransition measure is a continuous functional on the space of
2376: continual Young diagrams, and derived an explicit formula for the
2377: probability density of its $u$-coordinate. By substituting the level
2378: curve $\tilde{g}_\alpha$ in the formula from \cite{romik}, we will get
2379: exactly the semicircle density \eqref{eq:semicircle}, proving Theorem
2380: 2.
2381:
2382: Let $\lambda:\lambda(1)\ge\lambda(2)\ge \ldots \ge\lambda(m)>0$ be a
2383: Young diagram with $k=|\lambda|=\sum_i \lambda(i)$ cells. A cell
2384: $c=(i,j)\in\lambda$ \ $(1\le i\le m, 1\le j \le \lambda(i))$ is called
2385: a \emph{corner} cell if removing it leaves a Young diagram
2386: $\lambda\setminus c$, or in other words if $j=\lambda(i)$ and ($i=m$
2387: or $\lambda(i)>\lambda(i+1)$). If $T$ is a Young tableau of shape
2388: $\lambda$, let $c_{\textrm{max}}(T)$ be the cell containing the
2389: maximal entry $k$ in $T$. Obviously $c_{\textrm{max}}(T)$ is a corner
2390: cell of $\lambda$.
2391:
2392: The \emph{cotransition measure} of $\lambda$ is the probability
2393: measure $\mu_\lambda$ on corner cells of $\lambda$, which assigns to a
2394: corner cell $c$ measure
2395: \begin{equation}\label{eq:cotransition}
2396: \mu_\lambda(c) = \frac{d(\lambda \setminus c)}{d(\lambda)}
2397: \end{equation}
2398: (with $d(\lambda)$ as in \eqref{eq:hook}.) This is a probability
2399: measure, since one may divide up the $d(\lambda)$ tableaux of shape
2400: $\lambda$ according to the value of $c_{\textrm{max}}(T)$; for any
2401: corner cell $c$, there are precisely $d(\lambda\setminus c)$ tableaux
2402: for which $c_{\textrm{max}}(T)=c$. In other words $\mu_\lambda$ describes
2403: the distribution of $c_{\textrm{max}}(T)$, for a uniform random choice
2404: of a tableau $T$ of shape $\lambda$.
2405:
2406: It is fascinating that there exists a simple algorithm to sample from
2407: $\mu_\lambda$. This is known as the \emph{hook walk} algorithm of
2408: Greene-Nijenhuis-Wilf, and it can be described as follows: Choose a
2409: cell $c=(i,j)\in\lambda$ uniformly among all $k$ cells. Now execute a
2410: random walk, replacing at each step the cell $c$ with a new cell $c'$,
2411: where $c'$ is chosen uniformly among all cells which lie either to the
2412: right of, or (exclusive or) below $c$. The walk terminates when a corner
2413: cell is reached, and it can be shown \cite{greenenij} that the probability
2414: of reaching $c$ is given by \eqref{eq:cotransition}. Figure 6 shows a
2415: Young diagram, its corner cells and a sample hook walk path.
2416:
2417: \begin{figure}[h!]
2418: \begin{center}
2419: \begin{picture}(100,110)(0,0)
2420: \multiput(0,100)(0,-10){10}{\framebox(10,10)}
2421: \multiput(10,100)(0,-10){10}{\framebox(10,10)}
2422: \multiput(20,100)(0,-10){10}{\framebox(10,10)}
2423: \multiput(30,100)(0,-10){8}{\framebox(10,10)}
2424: \multiput(40,100)(0,-10){7}{\framebox(10,10)}
2425: \multiput(50,100)(0,-10){5}{\framebox(10,10)}
2426: \multiput(60,100)(0,-10){5}{\framebox(10,10)}
2427: \multiput(70,100)(0,-10){5}{\framebox(10,10)}
2428: \multiput(80,100)(0,-10){3}{\framebox(10,10)}
2429: \multiput(90,100)(0,-10){1}{\framebox(10,10)}
2430: \put(15,85){\circle{3}}
2431: \put(45,85){\circle{3}}
2432: \put(45,65){\circle{3}}
2433: \put(65,65){\circle{3}}
2434: \put(75,65){\circle{3}}
2435: \multiput(18,85)(3,0){9}{\circle*{1}}
2436: \multiput(45,81)(0,-3){5}{\circle*{1}}
2437: \multiput(49,65)(3,0){5}{\circle*{1}}
2438: \multiput(69,65)(3,0){2}{\circle*{1}}
2439: \multiput(22,11)(2,0){4}{\line(0,1){8}}
2440: \multiput(32,31)(2,0){4}{\line(0,1){8}}
2441: \multiput(42,41)(2,0){4}{\line(0,1){8}}
2442: \multiput(72,61)(6,0){2}{\line(0,1){8}}
2443: \multiput(74,61)(2,0){2}{\line(0,1){2}}
2444: \multiput(74,67)(2,0){2}{\line(0,1){2}}
2445: \multiput(82,81)(2,0){4}{\line(0,1){8}}
2446: \multiput(92,101)(2,0){4}{\line(0,1){8}}
2447: \end{picture}
2448: \caption{A Young diagram, its corners and a hook walk path}
2449: \end{center}
2450: \end{figure}
2451:
2452: Now consider a sequence
2453: $\lambda_n:\lambda_n(1)\ge\lambda_n(2)\ge\lambda_n(3)\ge \ldots $ of
2454: Young diagrams for which, under suitable scaling, the shape converges
2455: to some limiting shape described by a continuous function. More
2456: precisely, let $f_{\lambda_n}(x)$ be as in \eqref{eq:lambdagraph},
2457: and let $g_{\lambda_n}$ be its rotated coordinate version. Let
2458: $f_\infty:[0,\infty)\to[0,\infty)$ be a weakly decreasing function,
2459: and let $g_\infty$ be its rotated coordinate version, a 1-Lipschitz
2460: function. In this more general setting, think of $g_{\lambda_n}$ and
2461: $g_\infty$ as functions defined on all $\mathbb{R}$. Assume that:
2462: there exists an $M>0$ such that $f_\infty(x) = 0$ for $x\ge M$, and on
2463: $[0,M]$ $f$ is twice continuously differentiable, and its derivative
2464: is bounded away from $0$ and $\infty$ (equivalently, for some $K<0<K'$,
2465: $g_\infty(u)=|u|$ for $u \notin(K,K')$, and $g$ is twice continuously
2466: differentiable in $[K,K']$ with derivative bounded awaw from -1 and
2467: 1). Finally, assume that
2468: $$ || g_{\lambda_n} - g_\infty ||_\infty \xrightarrow[n\to\infty]{}
2469: 0. $$
2470: For any $n$, let $(I_n,J_n)$ be a
2471: $\mu_{\lambda_n}$-distributed random vector. Let $X_n = I_n/n,
2472: Y_n=J_n/n$. We paraphrase results from \cite{romik}.
2473:
2474: \paragraph{Theorem 10.} (Romik \cite{romik}, Theorems 1(b), 6) As
2475: $n\to\infty$, $(X_n,Y_n)$ converges in distribution to the random
2476: vector
2477: $$ (X,Y) := \left( \frac{V+U}{2}, \frac{V-U}{2} \right), $$ where
2478: $V=g_\infty(U)$ and $U$ is a random variable on $[K,K']$ with density function
2479: \begin{equation}\label{eq:cotransition2}
2480: \phi_U(x) = \frac{2}{\pi A} \cos\left( \frac{\pi
2481: g_\infty'(x)}{2}\right) \sqrt{(x-K)(K'-x)} \exp\left(\frac{1}{2}
2482: \int_K^{K'} \frac{g_\infty'(u)}{x-u}du \right),
2483: \end{equation}
2484: with
2485: $$ A = \int_0^M f_\infty(x)dx = \int_K^{K'} (g_\infty(u)-|u|)du $$
2486: and the integral in the exponential being a principal value integral.
2487:
2488: \paragraph{Proof of Theorem 2.} We may assume that $0<\alpha<1/2$. The
2489: proof of Theorem 2 now consists of an observation, a remark, and a
2490: computation.
2491:
2492: The observation is that the distribution of the location of the
2493: $k_n$-th entry in a random tableau $T \in {\cal T}_n$ is the
2494: distribution of the maximal entry in the shape $\lambda_T^{k_n}$ of
2495: the $k_n$-th subtableau of $T$. Because by Theorem 8, this shape
2496: (suitably rescaled and rotated) converges in probability to
2497: $\tilde{g}_\alpha$ (Theorem 2 assumes $k_n/n^2 \to \alpha$), we may
2498: apply Theorem 10 and conclude that Theorem 2 is true with a density for
2499: $U_\alpha$ given by taking $g_\infty = \tilde{g}_\alpha$, $A=\alpha$,
2500: $-K=K'=\sqrt{2\alpha(1-\alpha)}$ in \eqref{eq:cotransition2}.
2501:
2502: The remark is that the above is not quite true, since
2503: $\tilde{g}_\alpha$ does not satisfy the assumptions of Theorem 10! The
2504: problem is that
2505: $$ -\lim_{\epsilon \searrow 0}
2506: \tilde{g}_\alpha'(-\sqrt{2\alpha(1-\alpha)}+\epsilon) = \lim_{\epsilon
2507: \searrow 0} \tilde{g}_\alpha'(\sqrt{2\alpha(1-\alpha)}-\epsilon) = 1,
2508: $$
2509: so the derivative is not bounded away from -1 and 1. However, since
2510: this only happens near the two boundary points, going over the
2511: computations in \cite{romik} shows that this is not a problem, and the
2512: formula \eqref{eq:cotransition2} is still valid in this case
2513: \footnote{ Alternatively, one may use the less explicit formula (8)
2514: from \cite{romik}, which is valid even without the assumption that
2515: $g_\infty'$ is bounded away from $\pm 1$, to verify directly that
2516: \eqref{eq:semicircle} is the cotransition measure of
2517: $\tilde{g}_\alpha$. }.
2518:
2519: The computation is the verification that \eqref{eq:cotransition2}
2520: gives the semicircle distribution \eqref{eq:semicircle} under the
2521: above substitutions. We compute, using \eqref{eq:diffmultline} and the
2522: identity $\cos(\arct v) = (1+v^2)^{-1/2}$:
2523: \begin{eqnarray*}
2524: \frac{2}{\pi A} &=& \frac{2}{\pi \alpha} \\
2525: \sqrt{(x-K)(K'-x)} &=& \sqrt{2\alpha(1-\alpha)-x^2} \\
2526: \exp\left(\frac{1}{2}
2527: \int_K^{K'} \frac{\tilde{g}_\alpha(u)}{x-u}du \right) &=&
2528: \sqrt{\frac{\alpha}{1-\alpha}} \cdot
2529: \sqrt{\frac{\frac{1}{2}-x^2}{2\alpha(1-\alpha)-x^2}}, \\
2530: \cos\left( \frac{\pi \tilde{g}_\alpha'(x)}{2}\right) &=& \cos
2531: \left( \arct \frac{(1-2\alpha)x}{\sqrt{2\alpha(1-\alpha)-x^2}} \right)
2532: \\ &=&
2533: \left(1+\frac{(1-4\alpha(1-\alpha))x^2}{2\alpha(1-\alpha)-x^2}
2534: \right)^{-1/2} =
2535: \frac{\sqrt{2\alpha(1-\alpha)-x^2}}{2\sqrt{\alpha(1-\alpha)}
2536: \sqrt{\frac{1}{2}-x^2}}
2537: \end{eqnarray*}
2538: Multiplying the above expressions gives
2539: $$ \phi_U(x) = \frac{1}{\pi\alpha(1-\alpha)}
2540: \sqrt{2\alpha(1-\alpha)-x^2}, \qquad |x|\le \sqrt{2\alpha(1-\alpha)},
2541: $$
2542: as claimed.
2543: \qed
2544:
2545: \section{Plane partitions}
2546:
2547: We prove Theorem 4 on the limit shape of plane partitions of an
2548: integer $m$ over an $n\times n$ square diagram, when $n^6 = o(m)$. The
2549: basic observation relating this to Young tableaux is that in this
2550: asymptotic regime, almost all plane partitions have distinct parts. A
2551: plane partition with distinct parts can be completely described by
2552: separately giving the order structure on its parts -- a square Young
2553: tableau -- and an unordered list of the parts, which is simply a
2554: linear partition of $m$ into $n^2$ distinct parts. The structure of
2555: these linear partitions is described by a limit shape theorem due to
2556: Vershik and Yakubovich \cite{vershikyak1},
2557: \cite{vershikyak2}. Combining these results will give us our proof of
2558: Theorem 4.
2559:
2560: We will use a result of Erd\"os and Lehner on partitions into a fixed
2561: number of summands.
2562:
2563: \paragraph{Theorem 11. (Erd\"os-Lehner \cite{erdoslehner})}
2564: Let $p(m,k)$ denote the number of partitions of $m$ into $k$
2565: parts. Let $q(n,k)$ denote the number of partitions of $m$ into $k$
2566: \emph{distinct} parts. If $m$ and $k$ are sequences of integers that
2567: tend to infinity in such a way that $k^3=o(m)$, then
2568: $$ \frac{q(m,k)}{p(m,k)} \xrightarrow[\qquad]{} 1. $$
2569: In words, if $k^3=o(m)$, almost all partitions of $m$ into $k$ parts
2570: have no repeated parts.
2571:
2572: \paragraph{Proof.} This is a combination of Corollary 4.3 and Lemma
2573: 4.4 in \cite{erdoslehner}.
2574: \qed
2575:
2576: \bigskip
2577: For a Young diagram $\lambda$, denote by $p_\lambda(m)$ the number of
2578: plane partitions of $m$ of shape $\lambda$. Denote by $q_\lambda(m)$
2579: the number of plane partitions of $m$ of shape $\lambda$ with all
2580: parts distinct.
2581:
2582: \paragraph{Lemma 11.} Let $\lambda_m$ be a sequence of Young
2583: diagrams. Let $k_m = |\lambda_m|$. If $k_m^3 = o(m)$ as $m\to\infty$,
2584: then
2585: $$ \frac{q_{\lambda_m}(m)}{p_{\lambda_m}(m)}
2586: \xrightarrow[m\to\infty]{} 1. $$
2587: In words, if $k_m^3 = o(m)$, almost all plane partitions of shape
2588: $\lambda_m$ have no repeated parts.
2589:
2590: \paragraph{Proof.} If $\lambda$ is a Young diagram of size
2591: $k=|\lambda|$, a plane partition of $m$ of shape $\lambda$ is
2592: described by the order structure on its parts, and the unordered set
2593: of the parts. This gives the equation
2594: $$ q_\lambda(m) = d(\lambda) q(m,k). $$
2595: We claim that
2596: \begin{equation} \label{eq:ppineq}
2597: p_\lambda(m) \le d(\lambda) p(m,k).
2598: \end{equation}
2599: This will prove the claim, since then we will have
2600: $$ \frac{q(m,k_m)}{p(m,k_m)} \le
2601: \frac{q_{\lambda_m}(m)}{p_{\lambda_m}(m)} \le 1, $$ and the Lemma will
2602: follow from Theorem 11. To prove \eqref{eq:ppineq}, we define a
2603: mapping that assigns injectively to each plane partition
2604: $\pi=(p_{i,j})_{(i,j)\in \lambda}$ a pair $(T,\mu)$, where
2605: $T=(t_{i,j})_{(i,j)\in\lambda}$ is a Young tableau of shape $\lambda$,
2606: and $\mu:\mu(1)\ge \mu(2)\ge \ldots \ge\mu(k)$ is a partition of $m$ into
2607: $k$ parts. The mapping is defined as follows. Define a linear order
2608: ``$\prec$'' on the cells $(i,j)$ of $\lambda$, by stipulating that
2609: $$ (i,j) \prec (i',j') \iff
2610: p_{i,j}>p_{i',j'}\textrm{ or }
2611: \bigg[p_{i,j}=p_{i',j'}\textrm{ and }
2612: \big(i<i'\textrm{ or }
2613: (i=i'\textrm{ and }j<j')\big)\bigg].
2614: $$
2615: Let $(i_1,j_1) \prec (i_2,j_2) \prec \ldots \prec (i_k,j_k)$ be the cells
2616: of $\lambda$ sorted in this linear ordering, and set $t_{i_l,j_l}=l$
2617: and $\mu(l) = p_{i_l,j_l}$, \ $l=1,2,\ldots,k$.
2618:
2619: It is easy to verify that the mapping is injective and has the
2620: required range. See Figure 7 for an illustration.
2621: \qed
2622:
2623: \begin{figure}[h!]
2624: \begin{center}
2625: $\pi =$\
2626: \begin{tabular}{lllll}
2627: 7 & 7 & 6 & 5 & 2 \\
2628: 7 & 6 & 5 & 5 & \\
2629: 7 & 5 & 2 & & \\
2630: 6 & & & &
2631: \end{tabular}
2632: $\xrightarrow[\qquad]{}
2633: \begin{array}{l}
2634: T =\
2635: \begin{tabular}{lllll}
2636: 1 & 2 & 5 & 8 & 12 \\
2637: 3 & 6 & 9 & 10 & \\
2638: 4 & 11 & 13 & & \\
2639: 7 & & & &
2640: \end{tabular} \\ \ \\
2641: \mu = 7,\,7,\, 7,\,7,\,6,\,6,\,6,\,5,\,5,\,5,\,2,\,2.
2642: \end{array}$
2643: \caption{Illustration of the proof of Lemma 11}
2644: \end{center}
2645: \end{figure}
2646:
2647: \bigskip
2648: Next, we recall the Vershik-Yakubovich limit shape theorem for
2649: partitions of $m$ into $k$ distinct summands, when $k=o(\sqrt{m})$.
2650:
2651: \paragraph{Theorem 12. (Vershik-Yakubovich \cite{vershikyak1},
2652: \cite{vershikyak2})} Let $m=m_n$ and $k=k_n$ grow to infinity
2653: as a function of some parameter $n$, in such a way that
2654: $k=o(\sqrt{m})$. Let $\lambda_n: \lambda_n(1)> \lambda_n(2) >
2655: \ldots >\lambda_n(k)$ be a sequence of uniform random partitions of $m$
2656: into $k$ distinct parts. Then for any $t\ge 0,\ \epsilon>0$,
2657: $$
2658: \mathbb{P}\left( \left|\frac{1}{k}\# \{ 1 \le l \le k : \lambda_n(l) >
2659: \frac{m}{k} t \} - e^{-t}\right|>\epsilon \right)
2660: \xrightarrow[n\to\infty]{} 0.
2661: $$
2662: In words, the graph of the Young diagram of a uniform random partition
2663: of $m$ into $k$ distinct parts, when $k=o(\sqrt{m})$, will with high
2664: probability resemble the limit shape $e^{-t}$.\footnote{Actually, it
2665: is more correct to say that this is the graph of the conjugate
2666: partition $\lambda'$.}
2667:
2668: \paragraph{Proof of Theorem 4.} Let $\pi=(p_{i,j})_{i,j=1}^n$ be the
2669: random plane partition of $m$ over the square diagram
2670: $\square_n$. Since $n^6=o(m)$ and $|\square_n|=n^2$, by Lemma 11 we may
2671: assume that $\pi$ was chosen uniformly among all plane partitions of
2672: $m$ of shape $\square_n$ \emph{with all parts distinct}, since this is
2673: a set of probability close to $1$ in ${\cal P}_{n,m}$. Equivalently,
2674: by the remarks at the beginning of this section, we may assume that
2675: $\pi$ is selected by choosing independently a random Young tableau $T
2676: \in {\cal T}_n$ and a random partition
2677: $\mu:\mu(1)>\mu(2)>\ldots>\mu(n^2)$ of $m$ into $n^2$ distinct parts,
2678: then setting $p_{i,j} = \mu(t_{i,j})$.
2679:
2680: Fix $0\le x,y<1$. Let $\alpha=L(x,y)$. Let $i=\lfloor n x \rfloor+1$,
2681: $j=\lfloor n y \rfloor + 1$, and $\beta = t_{i,j}/n^2$.
2682:
2683: We need to show that $\tilde{S}_\pi(x,y) = (n^2/m)p_{i,j}$ is with
2684: high probability very close to $-\log \alpha$. Let $\epsilon>0$ be
2685: small. From Theorem 1(i), we know that with (asymptotically) high
2686: probability
2687: \begin{equation}\label{eq:betapp}
2688: |\beta-\alpha| < \epsilon.
2689: \end{equation}
2690: Now, apply Theorem 12 for the random partition $\mu$ with
2691: $t=-\log(\alpha-2\epsilon)$. This gives that with high probability
2692: $$ \left| \frac{1}{n^2}\# \left\{ 1 \le l \le n^2 : \mu(l) >
2693: \frac{m}{n^2}(-\log(\alpha-2\epsilon))\right\} - (\alpha-2\epsilon) \right| <
2694: \epsilon, $$
2695: or equivalently, since $\mu(1)>\mu(2)>\ldots>\mu(n^2)$,
2696: $$ \left| \frac{1}{n^2} \max\left\{ 1 \le l \le n^2 : \mu(l) >
2697: \frac{m}{n^2}(-\log(\alpha-2\epsilon))\right\} - (\alpha-2\epsilon) \right| <
2698: \epsilon. $$
2699: This implies in particular that
2700: $$ \max\left\{ 1 \le l \le n^2 : \mu(l) >
2701: \frac{m}{n^2}(-\log(\alpha-2\epsilon))\right\} <
2702: n^2(\alpha-2\epsilon+\epsilon) = (\alpha-\epsilon)n^2, $$
2703: hence, since by \eqref{eq:betapp}, $t_{i,j} > (\alpha-\epsilon)n^2$,
2704: $$ p_{i,j} = \mu(t_{i,j}) \le \frac{m}{n^2}(-\log(\alpha-2\epsilon)). $$
2705: Apply Theorem 12 again with
2706: $t=-\log(\alpha+2\epsilon)$. This gives that with high probability
2707: $$ \left| \frac{1}{n^2}\# \left\{ 1 \le l \le n^2 : \mu(l) >
2708: \frac{m}{n^2}(-\log(\alpha+2\epsilon))\right\} - (\alpha+2\epsilon) \right| <
2709: \epsilon, $$
2710: or equivalently
2711: $$ \left| \frac{1}{n^2} \max\left\{ 1 \le l \le n^2 : \mu(l) >
2712: \frac{m}{n^2}(-\log(\alpha+2\epsilon))\right\} - (\alpha+2\epsilon) \right| <
2713: \epsilon. $$
2714: In particular this gives that
2715: $$ \max\left\{ 1 \le l \le n^2 : \mu(l) >
2716: \frac{m}{n^2}(-\log(\alpha+2\epsilon))\right\} >
2717: n^2(\alpha+2\epsilon-\epsilon) = (\alpha+\epsilon)n^2, $$
2718: hence, since by \eqref{eq:betapp}, $t_{i,j} < (\alpha+\epsilon)n^2$,
2719: $$ p_{i,j} = \mu(t_{i,j}) > \frac{m}{n^2}(-\log(\alpha+2\epsilon)). $$
2720: We have shown that the event
2721: $$
2722: -\log(\alpha+2\epsilon) < \frac{n^2}{m} p_{i,j} \le -\log(\alpha-2
2723: \epsilon)
2724: $$
2725: holds with asymptotically high probability. Since $\epsilon$ was
2726: arbitrary the result follows.
2727: \qed
2728:
2729: \section{Computations for the rectangular case}
2730:
2731: The proof of Theorem 5 involves exactly the same ideas as the proof of
2732: Theorem 1, with some more computations, which we include here for
2733: completeness. The proof that Theorem 6 follows from Theorem 5 is
2734: completely identical to the proof in section 6 that Theorem 4 follows
2735: from Theorem 1.
2736:
2737: Fix $0<\theta\le 1$ and $0<\alpha<1$. Our starting point is the
2738: rotated-coordinate formulation of the variational problem whose
2739: solution will yield the $\alpha$-level curve of the limit surface
2740: $L_\theta$. The computations leading to this variational problem are
2741: obvious generalizations of the corresponding computations for the
2742: square case $\theta=1$, and are omitted.
2743:
2744: \paragraph{Variational problem - the rectangular case.} A function
2745: $h:[-\theta \sqrt{2}/2, \sqrt{2}/2]\to[0,\infty)$ is called
2746: $\alpha$-admissible if $h$ is $1$-Lipschitz, and satisfies
2747: \begin{eqnarray} \label{eq:admiss1}
2748: h(-\theta\sqrt{2}/2) &=& \theta\sqrt{2}/2, \\
2749: \label{eq:admiss2}
2750: h(\sqrt{2}/2) &=& \sqrt{2}/2, \\
2751: \label{eq:admiss3}
2752: \int_{-\theta\sqrt{2}/2}^{\sqrt{2}/2}(h(u)-|u|)du &=& \alpha \theta.
2753: \end{eqnarray}
2754: Find the unique $\alpha$-admissible $h$ that minimizes
2755: $$ J(h) = -\frac{1}{2} \int_{-\theta\sqrt{2}/2}^{\sqrt{2}/2}
2756: \int_{-\theta\sqrt{2}/2}^{\sqrt{2}/2} h'(s)h'(t)\log|s-t|ds\,dt.
2757: $$
2758:
2759: \bigskip
2760: To derive the minimizer, first consider the case when $\alpha$ is
2761: small. In that case, we make an assumption on the form of the
2762: minimizer similar to \eqref{eq:guess}, but with a non-symmetric
2763: interval $[-\beta_1(\alpha),\beta_2(\alpha)]$, where $\beta_1\in
2764: (0,\theta\sqrt{2}/2)$, $\beta_2\in(0,\sqrt{2}/2)$. That is, we assume
2765: that $h'$ has the form
2766: \begin{equation}\label{eq:guessrect}
2767: h'(s)\ \ \ \textrm{ is }\left\{ \begin{array}{ll}
2768: = -1, & \textrm{if }-\theta\sqrt{2}/2 < s < -\beta_1, \\
2769: \in (-1,1), & \textrm{if }-\beta_1<s<\beta_2, \\
2770: = +1, & \textrm{if }\beta_2<s<\sqrt{2}/2. \end{array}\right.
2771: \end{equation}
2772: Replace the conditions \eqref{eq:admiss1}, \eqref{eq:admiss2},
2773: \eqref{eq:admiss3} with the equivalent set of conditions
2774: \begin{eqnarray} \label{eq:adm1}
2775: h(-\theta\sqrt{2}/2) &=& \theta\sqrt{2}/2, \\
2776: \label{eq:adm2}
2777: \int_{-\theta\sqrt{2}/2}^{\sqrt{2}/2}h'(u)du &=& (1-\theta)\sqrt{2}/2,\\
2778: \label{eq:adm3}
2779: \int_{-\theta\sqrt{2}/2}^{\sqrt{2}/2}uh'(u)du +
2780: \frac{1+\theta^2}{4}&=&\theta \alpha.
2781: \end{eqnarray}
2782: (In the square case $\theta=1$ we did not impose the restriction
2783: \eqref{eq:adm2} on $h$ as we expected $h$ to be even, i.e. $h'$ to be
2784: odd, so that the condition \eqref{eq:adm2} would be met
2785: automatically. Not anymore in the rectangular case!) Then the
2786: counterpart to \eqref{eq:multline} is: for $s \in (-\beta_1,\beta_2)$,
2787: %
2788: %
2789: %
2790: %\newcommand{\sq}{\sqrt{2}/2}
2791: %\newcommand{\tsq}{\theta\sqrt{2}/2}
2792: %\newcommand{\bo}{\beta_1}
2793: %\newcommand{\bt}{\beta_2}
2794: %\newcommand{\gamo}{\gamma_1}
2795: %\newcommand{\gamt}{\gamma_2}
2796: %
2797: %
2798: %
2799: \begin{multline}\label{eq:rect1}
2800: -\int_{-\beta_1}^{\beta_2} h'(t)\log|s-t|dt - \lambda s - \mu \\
2801: + (s+\tsq)\log|s+\tsq|+(s-\tsq)\log|s-\tsq| \\
2802: - (s+\bo)\log|s+\bo|-(s-\bt)\log|s-\bt| \qquad\qquad\qquad\qquad \\
2803: +\bo-\bt+(1-\theta)\sq
2804: = 0. \qquad\qquad\qquad
2805: \end{multline}
2806: Here $\lambda$, $\mu$ are the Lagrangian multipliers dual to the
2807: constraints \eqref{eq:adm3} and \eqref{eq:adm2}
2808: respectively. Differentiating \eqref{eq:rect1} with respect to $s$
2809: gives
2810: \begin{equation}\label{eq:rect2}
2811: -\int_{-\bo}^{\bt} \frac{h'(t)}{s-t}dt = \lambda + \log
2812: \frac{s+\bo}{s+\tsq} + \log \frac{\bt-s}{\sq-s}, \qquad
2813: s\in(-\bo,\bt).
2814: \end{equation}
2815: Introduce $a=(\bo+\bt)/2$, $b=(\bt-\bo)/2$, and substitute $s=a\xi+b,
2816: t=a\eta+b$. The above equation becomes
2817: \begin{equation}\label{eq:rect3}
2818: -\int_{-1}^1 \frac{h'(a\eta+b)}{\xi-\eta}d\eta = \lambda +
2819: \log\frac{1+\xi}{\gamo+\xi} + \log\frac{1-\xi}{\gamt-\xi};
2820: \end{equation}
2821: here
2822: \begin{equation}\label{eq:gammadef}
2823: \gamo=\frac{\bt-\bo+\theta\sqrt{2}}{\bo+\bt}, \qquad
2824: \gamt=\frac{\bo-\bt+\sqrt{2}}{\bo+\bt},
2825: \end{equation}
2826: and it is easy to check that $\gamo,\gamt>1$. Applying Theorem 9 to
2827: \eqref{eq:rect3} and using Lemma 8, we obtain
2828: \begin{multline}\label{eq:hprimerect}
2829: h'(a\xi+b) =
2830: \frac{1}{\pi^2(1-\xi^2)^{1/2}}(\pi\lambda\xi+I(\xi,\gamo)-I(-\xi,\gamt))
2831: + \frac{c'}{(1-\xi^2)^{1/2}}
2832: \\ \quad
2833: = \frac{\xi}{\pi(1-\xi^2)^{1/2}}\left(
2834: \lambda-\log\left(\gamo+\sqrt{\gamo^2-1}\right)
2835: -\log\left(\gamt+\sqrt{\gamt^2-1}\right) \right) \\
2836: + \frac{2}{\pi}\left[
2837: \arct\sqrt{\frac{(1+\xi)(\gamt-1)}{(1-\xi)(\gamt+1)}}
2838: - \arct\sqrt{\frac{(1-\xi)(\gamo-1)}{(1+\xi)(\gamo+1)}} \right]
2839: + c(1-\xi^2)^{-1/2},
2840: \end{multline}
2841: $c', c$ being arbitrary constants. As in the symmetric case, if $h'(s)$
2842: is to be bounded for $s\in(-\bo,\bt)$ (i.e. for $\xi\in(-1,1)$),
2843: necessarily
2844: \begin{equation}\label{eq:lambdarect}
2845: c=0, \qquad
2846: \lambda = \log\left(\gamo+\sqrt{\gamo^2-1}\right)
2847: +\log\left(\gamt+\sqrt{\gamt^2-1}\right).
2848: \end{equation}
2849: So we have
2850: \begin{equation}\label{eq:hpr}
2851: h'(a\xi+b) =
2852: \frac{2}{\pi}\left[
2853: \arct\sqrt{\frac{(1+\xi)(\gamt-1)}{(1-\xi)(\gamt+1)}}
2854: - \arct\sqrt{\frac{(1-\xi)(\gamo-1)}{(1+\xi)(\gamo+1)}} \right],
2855: \end{equation}
2856: for which indeed $|h'(a\xi+b)|\le 1$ holds.
2857:
2858: We still need to find $\bo$ and $\bt$. Using \eqref{eq:guessrect},
2859: rewrite \eqref{eq:adm2} and \eqref{eq:adm3} as, respectively,
2860: \begin{eqnarray}
2861: \label{eq:adm2new}
2862: \int_{-\bo}^{\bt} h'(t)dt &=& \bt-\bo, \\
2863: \label{eq:adm3new}
2864: \int_{-\bo}^{\bt} th'(t)dt + \frac{1}{2}(\bo^2+\bt^2) &=&
2865: \theta\alpha.
2866: \end{eqnarray}
2867: Now evaluating these integrals using \eqref{eq:hpr}, this gives the
2868: equations
2869: \begin{eqnarray*}
2870: a\left[\left( \sqrt{\gamt^2-1}-\gamt\right)-
2871: \left( \sqrt{\gamo^2-1}-\gamo\right)\right] &=& \bt-\bo, \\
2872: -\frac{a^2}{2}\sum_{i=1}^2 \left[1-\gamma_i^2+\gamma_i
2873: \sqrt{\gamma_i^2-1}\right] - b(\bt-\bo)+\frac{1}{2}(\bo^2+\bt^2) &=&
2874: \theta\alpha,
2875: \end{eqnarray*}
2876: ($a=(\bo+\bt)/2$, $b=(\bt-\bo)/2$). Excluding $\bo, \bt$ via
2877: \eqref{eq:gammadef}, we obtain two equations for $\gamo, \gamt$,
2878: namely
2879: \begin{eqnarray}
2880: \label{eq:eqgam1}
2881: \gamo-\theta\gamt = \frac{1+\theta}{2}\left[
2882: ((\gamt^2-1)^{1/2}-\gamt)-((\gamo^2-1)^{1/2}-\gamo) \right], \\
2883: \label{eq:eqgam2}
2884: \theta\alpha =
2885: \frac{(1+\theta)^2-(\gamo-\theta\gamt)^2}{2(\gamo+\gamt)^2} -
2886: \frac{(1+\theta)^2}{4(\gamo+\gamt)^2}
2887: \sum_{i=1}^2 \left[1-\gamma_i^2+\gamma_i
2888: \sqrt{\gamma_i^2-1}\right].\
2889: \end{eqnarray}
2890: It seems a minor miracle that these equations can be solved
2891: explicitly. Here is how. Isolating the difference of the radicals in
2892: the first equation, multiplying both sides of the resulting equation
2893: by the sum of radicals and cancelling the common factor $\gamo+\gamt$,
2894: we obtain
2895: $$
2896: \sqrt{\gamo^2-1}+\sqrt{\gamt^2-1}=\frac{1+\theta}{1-\theta}(\gamt-\gamo).
2897: $$
2898: (In particular, $\gamt>\gamo$.) Combining this with the initial
2899: equation, we express the radicals as linear combinations of $\gamo,
2900: \gamt$:
2901: \begin{eqnarray}\label{eq:radical1}
2902: \sqrt{\gamo^2-1}&=&-\frac{1+\theta^2}{1-\theta^2}\gamo +
2903: \frac{2\theta}{1-\theta^2}\gamt, \\
2904: \label{eq:radical2}
2905: \sqrt{\gamt^2-1}&=&-\frac{2\theta}{1-\theta^2}\gamo +
2906: \frac{1+\theta^2}{1-\theta^2}\gamt.
2907: \end{eqnarray}
2908: Plugging these expressions for the radicals into \eqref{eq:eqgam2},
2909: after collecting like terms, we obtain a \emph{quadratic} equation for
2910: $x=\gamt/\gamo$:
2911: $$
2912: x^2(\theta(1-\alpha)+\alpha)-x(1-\theta)(1-2\alpha)-(1-\alpha+\theta\alpha)
2913: = 0.$$
2914: Consequently
2915: $$ \frac{\gamt}{\gamo} = x =
2916: \frac{\theta\alpha+1-\alpha}{\alpha+\theta(1-\alpha)}, $$
2917: which, for $\alpha<1/2$, exceeds 1. Squaring both sides of
2918: \eqref{eq:radical1} and substituting $\gamt=x\gamo$, we solve for
2919: $\gamo$ to obtain
2920: \begin{equation} \label{eq:eqgamnew}
2921: \gamo =
2922: \frac{\alpha+\theta(1-\alpha)}{2\sqrt{\theta\alpha(1-\alpha)}},
2923: \qquad
2924: \gamt = \frac{\theta\alpha+1-\alpha}{2\sqrt{\theta\alpha(1-\alpha)}}.
2925: \end{equation}
2926: Direct checking reveals that these $\gamo, \gamt$ satisfy the
2927: equations \eqref{eq:radical1}, \eqref{eq:radical2} themselves as long as
2928: \begin{equation}\label{eq:condalphastar}
2929: \alpha \le \alpha^* := \frac{\theta}{1+\theta}.
2930: \end{equation}
2931: For $\alpha>\alpha^*$, the gammas do not satisfy
2932: \eqref{eq:radical1}. More precisely, $\gamo, \gamt$ would have satisfied
2933: this equation, had we considered the negative value of
2934: $\sqrt{\gamo^2-1}$. However, we need the positive value only. The
2935: source of the trouble here is that $\gamo=1$ for
2936: $\alpha=\alpha^*$. Tellingly, the boundary point $(-\bo,h(-\bo))$
2937: reaches the corner $(-\tsq,\tsq)$ of the rotated rectangle at
2938: $\alpha=\alpha^*$. Using \eqref{eq:gammadef}, we obtain: for $\alpha
2939: \le \alpha^*$,
2940: \begin{multline}\label{eq:betabeta}
2941: \qquad\qquad\qquad\qquad\quad\ \
2942: \bo = \sqrt{2\theta\alpha(1-\alpha)} - \alpha(1-\theta)\sq, \\
2943: \bt = \sqrt{2\theta\alpha(1-\alpha)} + \alpha(1-\theta)\sq.
2944: \qquad\qquad\qquad\qquad\qquad
2945: \end{multline}
2946: Using \eqref{eq:eqgamnew} and \eqref{eq:betabeta}, we simplify
2947: \eqref{eq:hpr} to
2948: \begin{equation}\label{eq:hprnew}
2949: h'(t) = \frac{2}{\pi}\arct \left[
2950: \frac{(1-\theta)\sqrt{\alpha(1-\alpha)} + \xi\sqrt{\theta}(1-2\alpha)}
2951: {\sqrt{\theta(1-\xi^2)}}\right],
2952: \end{equation}
2953: where
2954: $$ \xi = \frac{t-b}{a} =
2955: \frac{t-\alpha(1-\theta)\sq}{\sqrt{2\theta\alpha(1-\alpha)}}, \qquad
2956: t\in [-\beta_1,\beta_2].
2957: $$
2958: Furthermore, using
2959: $$ h(s) = \bo + \int_{-\bo}^s h'(t)dt = \bo + a\int_{-1}^\xi
2960: h'(a\eta+b) d\eta, $$
2961: \eqref{eq:hpr}, and \eqref{eq:eqx4}, we obtain
2962: \begin{multline}\label{eq:formulah}
2963: h(s) = h(a\xi+b) = \bo \\ +\frac{2a}{\pi} \bigg[ -(\xi+\gamo) \arct
2964: \sqrt{\frac{(1-\xi)(\gamo-1)}{(1+\xi)(\gamo+1)}} + (\xi-\gamt) \arct
2965: \sqrt{\frac{(1+\xi)(\gamt-1)}{(1-\xi)(\gamt+1)}} \\
2966: + \frac{1}{2}\left(\arcs \xi+\frac{\pi}{2}\right)
2967: \left(\sqrt{\gamt^2-1}-\sqrt{\gamo^2-1}\right) +
2968: \frac{\pi}{2}(\gamo-1) \bigg].
2969: \end{multline}
2970: Finally, combining \eqref{eq:lambdarect} and \eqref{eq:eqgamnew}, we
2971: compute
2972: \begin{equation}\label{eq:lambdanew}
2973: \lambda = \log \frac{\theta(1-\alpha)}{\sqrt{\theta\alpha(1-\alpha)}}
2974: + \log \frac{(1-\alpha)}{\sqrt{\theta\alpha(1-\alpha)}} =
2975: \log\frac{1-\alpha}{\alpha},
2976: \end{equation}
2977: the same value as in the square case!
2978:
2979: It remains to consider the range $\alpha^* < \alpha \le 1/2$. In this
2980: case, it turns out that the formulas for the parameters $\bo, \bt$
2981: remain the same, while the formula for the corresponding $h'$ changes
2982: slightly. What is different is that now $h'(s)=1$ for $s\in [-\tsq,
2983: -\bo]$ and
2984: $$ h(-\bo) = \theta\sqrt{2} - \bo. $$
2985: The latter condition means that now the boundary point $[-\bo,
2986: h(-\bo)]$ lies on the longer side of the rectangle. The starting point
2987: now is a modification of \eqref{eq:rect2}, stemming from
2988: $h'(s)\equiv 1$, rather than $-1$, for $s\in [-\tsq,-\bo]$, namely
2989: $$
2990: -\int_{-1}^1 \frac{h'(a\eta+b)}{\xi-\eta}d\eta = \lambda - \log
2991: \frac{1+\xi}{\gamo+\xi} + \log \frac{1-\xi}{\gamt-\xi}.
2992: $$
2993: This leads to
2994: $$ h'(a\xi+b) =
2995: \frac{1}{\pi^2(1-\xi^2)^{1/2}}(\pi\lambda\xi-I(\xi,\gamo) -
2996: I(-\xi,\gamt)) + \frac{c'}{(1-\xi^2)^{1/2}}, $$
2997: ($c'$ being a constant),
2998: compare with the first line in \eqref{eq:hprimerect}, whence to
2999: $$
3000: \lambda = \log\left(\gamt+\sqrt{\gamt^2-1}\right)
3001: -\log\left(\gamo+\sqrt{\gamo^2-1}\right),
3002: $$
3003: compare with \eqref{eq:lambdarect}, and
3004: \begin{equation*}\label{eq:hpr2}
3005: h'(a\xi+b) =
3006: \frac{2}{\pi}\left[
3007: \arct\sqrt{\frac{(1-\xi)(\gamo-1)}{(1+\xi)(\gamo+1)}}
3008: + \arct\sqrt{\frac{(1+\xi)(\gamt-1)}{(1-\xi)(\gamt+1)}} \right],
3009: \end{equation*}
3010: compare with \eqref{eq:hpr}. After integration, the final formula for
3011: $h(s)$ is
3012: \begin{multline}\label{eq:formulah2}
3013: h(s) = \theta\sqrt{2} - \bo \\ +\frac{2a}{\pi} \bigg[ (\xi+\gamo) \arct
3014: \sqrt{\frac{(1-\xi)(\gamo-1)}{(1+\xi)(\gamo+1)}} + (\xi-\gamt) \arct
3015: \sqrt{\frac{(1+\xi)(\gamt-1)}{(1-\xi)(\gamt+1)}} \\
3016: + \frac{1}{2}\left(\arcs \xi+\frac{\pi}{2}\right)
3017: \left(\sqrt{\gamo^2-1}+\sqrt{\gamt^2-1}\right) +
3018: \frac{\pi}{2}(1-\gamo) \bigg].
3019: \end{multline}
3020: We add that,
3021: despite the difference between the two formulas for $\lambda$ -- the
3022: one above for $\alpha \ge \alpha^*$ and \eqref{eq:lambdarect} for
3023: $\alpha\le \alpha^*$ -- the eventual expression is still that in
3024: \eqref{eq:lambdanew}. The ``secret'' is that
3025: $$ \sqrt{\gamo^2-1} =
3026: \frac{|\sqrt{\alpha}-\sqrt{\theta(1-\alpha)}|}{2\sqrt{\theta\alpha(1-\alpha)}},
3027: $$
3028: with $\sqrt{\alpha}-\sqrt{\theta(1-\alpha)}$ changing its sign at
3029: $\alpha^*$.
3030:
3031: It remains to prove that $h$ is indeed a minimizer. Let $\alpha <
3032: \alpha^*$. Consider $ s \in [-\tsq,-\bo]$. Since $h'(s)=-1$, we need
3033: to show that $F(s,\alpha)\ge 0$, where $F(s,\alpha)$ is the left-hand
3034: side expression in \eqref{eq:rect1}. The above computations show that
3035: $F(s,\alpha) \equiv 0$ for $s\in (-\bo,\bt)$. As in the square case,
3036: for fixed $s\in [-\tsq,\bo]$ let $\hat{\alpha}\in (\alpha,\alpha^*)$
3037: be defined by $s=-\bo(\hat{\alpha})$. Then $F(s,\hat{\alpha})=0$,
3038: and we will prove $F(s,\alpha)\ge 0$ if we show that $\partial
3039: F(s,x)/\partial x < 0$ for all $x\in [\alpha,\hat{\alpha}]$. Since
3040: $0\in (-\bo(x),\bt(x))$, $F(0,x)=0$, and we use the latter equation to
3041: exclude the Lagrangian multiplier $\mu$ in the expression for
3042: $F(s,x)$. Then, an easy computation shows that
3043: \begin{equation}\label{eq:easycomp}
3044: \frac{\partial F(s,x)}{\partial x} = -\int_{-\bo}^{\bt} \frac{\partial
3045: h_x(t)}{\partial x} \log|s-t|dt + \int_{-\bo}^{\bt} \frac{\partial
3046: h_x(t)}{\partial x} \log|t|dt -s \frac{d\lambda}{dx},
3047: \end{equation}
3048: where $\beta_i = \beta_i(x), \lambda=\lambda(x)$ are given by
3049: \eqref{eq:betabeta} and \eqref{eq:lambdanew}. Let us evaluate
3050: $\partial h_x(s)/\partial x$ for $s\in (-\bo,\bt)$. Differentiating
3051: \eqref{eq:rect2} with respect to $x$ we obtain
3052: $$ -\int_{-\bo}^{\bt} \frac{\partial h_x(t)/\partial x}{s-t}dt =
3053: \frac{d\lambda}{dx} = - \frac{1}{x(1-x)}.
3054: $$
3055: Then, using Theorem 9 and \eqref{eq:lambdaterm},
3056: \begin{equation}\label{eq:using}
3057: \frac{\partial h_x(s)}{\partial x} = -\frac{\xi}{\pi
3058: x(1-x)(1-\xi^2)^{1/2}} + \frac{c}{(1-\xi^2)^{1/2}}.
3059: \end{equation}
3060: Here we have to set $c=0$, as the equation \eqref{eq:adm2new} -- upon
3061: differentiation with respect to $x$ -- leads to
3062: \begin{equation}\label{eq:leadsto}
3063: \int_{-1}^1 \frac{\partial h_x(t)}{\partial x} d\xi = 0.
3064: \end{equation}
3065: Hence
3066: \begin{equation}\label{eq:hence}
3067: \frac{\partial h_x(s)}{\partial x} = -\frac{\xi}{\pi
3068: x(1-x)(1-\xi^2)^{1/2}}.
3069: \end{equation}
3070: Plugging this expression into \eqref{eq:easycomp}, integrating by
3071: parts, and using \eqref{eq:unnumbered}, we transform
3072: \eqref{eq:easycomp} into
3073: \begin{equation}\label{eq:transform}
3074: \frac{\partial F(s,x)}{\partial x} =
3075: -\frac{\sqrt{(s-b)^2-a^2}}{x(1-x)} < 0.
3076: \end{equation}
3077: Let $\alpha \in (\alpha^*,1/2]$, and $s<-\bo(\alpha)$ again. Since now
3078: $h'(s)=1$, we need to show that $F(s,\alpha)\le 0$. Let
3079: $\tilde{\alpha} \in (\alpha^*,\alpha)$ be defined by $-s =
3080: \bo(\tilde{\alpha})$. ($\tilde{\alpha}$ exists, uniquely, because
3081: $\bo(x)$ is decreasing on $(\alpha,\alpha^*)$ and $-s < \bo(\alpha)$.)
3082: Then $F(s,\tilde{\alpha})=0$, and so again we need to show that
3083: $\partial F(s,x)/\partial x < 0$. The formula \eqref{eq:using}
3084: continues to hold, and so does \eqref{eq:leadsto}, since now we have
3085: $$ \int_{-\bo}^{\bt} h'(t)dt = \bt + \bo - \tsq, $$
3086: and $h'(-\bo)=1$. Therefore \eqref{eq:hence} remains valid, which
3087: implies \eqref{eq:transform}.
3088:
3089: Analogously, $F(s,\alpha)\le 0$ for $s\ge \bt(\alpha)$ and $\alpha \in
3090: (0,1/2]$. This finishes the proof that $h$ is the minimizer, the claim
3091: which forms the core of the proof of Theorem 5.
3092: \qed
3093:
3094: \section{Discussion}
3095:
3096: \subsection{Plancherel measure}
3097:
3098: Let ${\cal Y}_k$ denote the set of Young diagrams of area $k$. The
3099: Plancherel measures are the family of probability measures $\mu_k$ on
3100: ${\cal Y}_k$, defined by
3101: \begin{equation}\label{eq:plancherel}
3102: \mu_k(\lambda) = \frac{d(\lambda)^2}{k!}, \qquad (\lambda \in {\cal Y}_k).
3103: \end{equation}
3104: Alternatively, $\mu_k$ is sometimes defined as a measure on all Young
3105: tableaux of size $k$, where
3106: $$ \mu_k(T) = \frac{d(\textrm{shape}(T))}{k!}. $$ The measure on
3107: diagrams is then the projection of the measure on tableaux under the
3108: mapping that assigns to each tableau its shape. The measures $\mu_k$
3109: are a \emph{projective family} of measures, in the following sense: If
3110: $T$ is a $\mu_k$-random tableau, then the tableau $T'$ of size $k-1$
3111: obtained by deleting the $k$-th entry from $T$ is a $\mu_{k-1}$-random
3112: tableau. Therefore, all the $\mu_k$'s can be encompassed by a single
3113: object $\mathbb{P}$, the \emph{infinite Plancherel measure}, which is
3114: a measure on infinite tableaux -- i.e. fillings of the squares in the
3115: positive quadrant of the plane with the positive integers that are
3116: increasing along rows and columns -- for which the marginal
3117: distribution of the shape of the $k$-th subtableau (the set of squares
3118: where the entry of the infinite tableau is $\le k$) is given by
3119: \eqref{eq:plancherel}. In other words, $\mathbb{P}$ can be thought of
3120: as a measure on all sequences $\emptyset = \lambda_0 \subset \lambda_1
3121: \subset \lambda_2 \subset ... $ of Young diagrams, where $\lambda_k$
3122: has size $k$ and is obtained from $\lambda_{k-1}$ by the addition of a
3123: box. So $\mathbb{P}$ is simply a natural Markovian coupling of the
3124: measures \eqref{eq:plancherel}, known sometimes as the
3125: \emph{Plancherel growth process}.
3126:
3127: Much is known about Plancherel measure. It arises naturally in
3128: representation theory, as a natural measure on the irreducible
3129: representations of the symmetric group, and in combinatorics, as the
3130: distribution of the output of the RSK algorithm applied to a uniform
3131: random permutation in $S_k$. In particular, the length of the first
3132: row $\lambda_k(1)$ of a $\mu_k$-random Young diagram has the same
3133: distribution as the length $l_n(\pi)$ of the longest increasing
3134: subsequence of a uniform random permutation $\pi$ in $S_k$, an important
3135: permutation statistic.
3136:
3137: Logan-Shepp \cite{loganshepp} and Vershik-Kerov \cite{vershikkerov1},
3138: \cite{vershikkerov2} proved that the graph of a $\mu_k$-random Young
3139: diagram, when rescaled by a factor of $\sqrt{k}$ along each axis and
3140: drawn in rotated coordinates, with high probability resembles the
3141: limit shape
3142: $$ \Omega(u) = \left\{ \begin{array}{ll}
3143: \frac{2}{\pi}(u\arcs(u/\sqrt{2})+\sqrt{2-u^2}) & |u| \le \sqrt{2}, \\
3144: |u| & |u| > \sqrt{2}, \end{array} \right.$$
3145: see Figure 8.
3146:
3147: %\vspace{-10.0 pt}
3148: \begin{figure}[h!]
3149: \begin{center}
3150: \includegraphics{vk.eps}
3151: \caption{The limit shape $v=\Omega(u)$}
3152: \end{center}
3153: \end{figure}
3154: %\vspace{-178.0 pt} \quad\quad\quad
3155: %\quad\quad\qquad\qquad\qquad\qquad\qquad {\footnotesize $v=\Omega(u)$}
3156: %
3157: %\vspace{160.0 pt}
3158: Gribov \cite{gribov} noted that this can be reinterpreted as a theorem
3159: on the limit \emph{surface} of the Plancherel-random tableau, much in
3160: the same spirit as Theorem 1. If $T$ is a $\mu_k$-random Young
3161: tableau, then after rescaling the graph of $T$ is approximately
3162: described in rotated coordinates by the surface $\Sigma:D\to[0,1]$,
3163: where
3164: $$ D = \{ (u,v): |u| \le \sqrt{2},\ \, |u| \le v \le \Omega(u) \} $$
3165: is the two dimensional domain bounded between the graphs $|u|$ and
3166: $\Omega(u)$, and for each $0 < \alpha < 1$, the $\alpha$-level curve
3167: of $\Sigma$ is
3168: \begin{equation*}\label{eq:time}
3169: \{ (u,v)\in D : |u| \le \sqrt{2 \alpha},\ \, v =
3170: \sqrt{\alpha}\Omega(u/\sqrt{\alpha}) \},
3171: \end{equation*}
3172: a shrunken copy of $\Omega$.
3173:
3174: The approach in the papers of Logan-Shepp and Vershik-Kerov was the
3175: variational approach, of analyzing the limiting integral functional
3176: arising from \eqref{eq:plancherel}. Kerov \cite{kerov} considered the
3177: following more dynamical approach: Assume that we have selected the
3178: $\mu_k$-random diagram $\lambda_k$. Since under $\mathbb{P}$, the
3179: sequence $\lambda_1 \subset \lambda_2 \subset ... $ is a
3180: (nonhomogeneous) Markov chain with values in the Young graph, there is
3181: a measure $\nu$ on the exterior corners of $\lambda_k$ (the boxes in
3182: the complement of $\lambda_k$ that can be added to $\lambda_k$ to form
3183: a Young diagram of size $k+1$), such that if we choose a $\nu$-random
3184: corner of $\lambda_k$ and add the new box there, the resulting diagram
3185: $\lambda_{k+1}$ will have distribution $\mu_{k+1}$. In other words,
3186: $\nu$ is the probability transition measure of the Markov chain
3187: $(\lambda_k)$. It is known as the \emph{transition measure of the
3188: diagram $\lambda_k$}, and is in a sense dual to the co-transition
3189: measure discussed in section 5.
3190:
3191: Kerov showed that in the limit when the graph of the diagram
3192: $\lambda_k$ becomes a smooth curve, the transition measure converges
3193: to a limit. Imagine that in the limit, instead of attaching a new box
3194: at a $\nu$-random corner, one attaches a $\nu$-fraction of a box at
3195: each corner. So the curve grows in the ``tangent'' direction given by
3196: $\nu$. Thus, the Plancherel growth process can be described in the
3197: limit as a smooth flow on the (infinite-dimensional) space of
3198: shapes. Kerov showed that $\Omega(u)$ is the unique shape which, after
3199: rescaling, is invariant under this flow, and that this fixed point is
3200: an attractor of the flow; this explains, in a way, (though does not
3201: formally prove) its appearance as the limit shape for
3202: Plancherel-random diagrams. Remarkably, the transition measure of
3203: $\Omega$ (the limiting direction of the flow) is the semicircle
3204: distribution.
3205:
3206: Another interesting direction stemming from the study of Plancherel
3207: measure is the connection to longest increasing subsequences of random
3208: permutations. The limit shape result of Logan-Shepp and Vershik-Kerov
3209: implies that the length $l_n(\pi)$ of the longest increasing
3210: subsequence of a random permutation $\pi\in S_n$ is with high
3211: probability at least $(1-o(1))2 \sqrt{n}$. Using additional arguments
3212: (which were an inspiration for our proof of Theorem 3), Vershik and
3213: Kerov showed also that $l_n(\pi)$ is with high probability at most
3214: $(1+o(1))2\sqrt{n}$, solving the so-called Ulam's problem. More
3215: recently, Baik, Deift and Johansson \cite{baiketal} showed that the
3216: fluctuation of $l_n(\pi)$ around its asymptotic value $2\sqrt{n}$ has
3217: a limiting distribution. More precisely,
3218: $$ \frac{l_n(\pi) - 2\sqrt{n}}{n^{1/6}}
3219: \xrightarrow[n\to\infty]{\textrm{ in distribution }} F. $$
3220: Here $F$ is the Tracy-Widom distribution from random matrix theory,
3221: defined as
3222: $$ F(t) = \exp\left(-\int_t^\infty (x-t)u(x)^2 dx \right), $$
3223: where $u(x)$ is the solution of the Painlev\'e II equation
3224: $u''(x)=2u(x)^3 + xu(x)$ that is asymptotic to the Airy function
3225: $\textrm{Ai}(x)$ as $x\to\infty$.
3226: Other results along those lines can be found in \cite{baiketal2},
3227: \cite{borodinetal}, \cite{johansson}, \cite{okounkov}; see also the
3228: survey \cite{aldousdiaconis}
3229:
3230: The distribution $F$ appears in random matrix theory as the limiting
3231: distribution of the maximal eigenvalue of a GUE random
3232: matrix. Following the Baik-Deift-Johansson result, it was found that
3233: there are many striking parallels between Plancherel measure and
3234: random matrix ensembles, see \cite{borodinetal}, \cite{oconnell},
3235: \cite{ivanovolsh}. In particular, the transition measure of the
3236: Plancherel-random diagram converges to the semicircle law, which is
3237: also the limiting distribution of the empirical eigenvalue
3238: distribution in the GUE and GOE random matrix ensembles. Ivanov and
3239: Olshanski \cite{ivanovolsh} showed that this similarity is no mere
3240: coincidence, but in fact appears also in the finer fluctuations of the
3241: transition measure and eigenvalue distribution measure around the
3242: semicircle distribution.
3243:
3244: \subsection{The random square tableau as a deformation of Plancherel
3245: measure}
3246:
3247: The reader familiar with the works of Logan-Shepp and Vershik-Kerov
3248: will undoubtedly have noticed the similarity between these results and
3249: our analysis of the square tableau model. Define for each positive
3250: integer $n$ and each $1 \le k \le n^2$, the probability measure
3251: $\nu_{n,k}$ on ${\cal Y}_k$, by
3252: \begin{equation}\label{eq:deform}
3253: \nu_{n,k}(\lambda) = \frac{d(\lambda)d(\square_n\setminus
3254: \lambda)}{d(\square_n)}, \qquad (\lambda \in {\cal Y}_k),
3255: \end{equation}
3256: where $d(\square_n\setminus\lambda)$ is taken as $0$ if $\lambda
3257: \nsubseteq \square_n$. The measure $\nu_{n,k}$ is the distribution of
3258: the $k$-th subtableau of a random $n\times n$ square tableau, and our
3259: entire approach revolved around the analysis of its properties. It is
3260: remarkable how many of the ideas used in the study of Plancherel
3261: measure we have found useful in our study of square tableaux; first,
3262: and most obviously, the variational problem that arises from
3263: \eqref{eq:deform} resembles the variational problem studied by
3264: Logan-Shepp and Vershik-Kerov. Although our approach in solving the
3265: variational problem relied on the more methodical use of the inversion
3266: formula for Hilbert transforms (an approach that could be applied the
3267: Plancherel case as well!), we were greatly inspired by the methods
3268: used in the Plancherel case. Secondly, our proof of Theorem 3 and the
3269: treatment of the boundary of the square also follows closely the ideas
3270: of Vershik and Kerov (with the notable difference, that our proof of
3271: the upper bound uses the lower bound!). Finally, our Theorem 2 on the
3272: location of particular entries, was inspired by Kerov's differential
3273: model \cite{kerov} for Plancherel growth. By postulating the existence
3274: of an analogous differential growth model for the $k$-subtableaux of
3275: the square tableau, we were able to guess Theorem 2 from the formula
3276: \eqref{eq:partialdiff}. This was later verified by a different method,
3277: using the result from \cite{romik}.
3278:
3279: Take another look at \eqref{eq:deform} and \eqref{eq:plancherel}. The
3280: defining equations for $\mu_k$ and $\nu_{n,k}$ seem superficially
3281: similar at best. In fact, they are closely related, and when $k$ is
3282: very small these measures are quite close. To make this precise, we
3283: first note the following curious identity. Define the \emph{falling
3284: power} $a^{\downarrow b}$ of $a$ as $a^{\downarrow
3285: b}=a(a-1)(a-2)\ldots (a-b+1)$. Then:
3286:
3287: \paragraph{Lemma 12.} If $\lambda \in {\cal Y}_k$, $\lambda \subset
3288: \square_n$, then
3289: $$
3290: \frac{\nu_{n,k}(\lambda)}{\mu_k(\lambda)} =
3291: \frac{\prod_{j=1}^{\lambda'(1)} (n+j-1)^{\downarrow j}
3292: \cdot \prod_{j=1}^{\lambda(1)} (n+j-1)^{\downarrow j}}
3293: {(n^2)^{\downarrow k}}
3294: $$
3295:
3296: \paragraph{Proof.} Use the hook formula \eqref{eq:hook}. A computation
3297: similar to the one in the proof of \eqref{eq:amus} shows that many of
3298: the terms cancel. We omit the relatively simple details.
3299: \qed
3300:
3301: \bigskip
3302: It follows using elementary estimates, which we again omit for the
3303: sake of brevity, that
3304:
3305: \paragraph{Theorem 13.} If $n\to \infty$ and $k=k(n)$ is such that
3306: $k=o(n^{2/3})$, then
3307: $$ \frac{\nu_{n,k}(\lambda)}{\mu_k(\lambda)}
3308: \xrightarrow[n\to\infty]{} 1 $$
3309: uniformly on the support ${\cal Y}_{n,k}$ of $\nu_{n,k}$ (the set of
3310: diagrams of size $k$ contained in $\square_n$). In particular, the
3311: total variation distance
$$ ||\nu_{n,k}-\mu_k||_1 := \sum_{\lambda \in {\cal Y}_k}
|\nu_{n,k}(\lambda) - \mu_k(\lambda)| \xrightarrow[n\to\infty]{} 0. $$
\vspace{-25.0 pt}
\qed
\bigskip
So in fact, when $k$ is small, $\nu_{n,k}$ is a kind of deformation of
the Plancherel measure $\mu_k$. In particular, for $k$ fixed and $n$
going to infinity, this implies the not-entirely-trivial fact that
$\mu_k$ is a probability measure. We remark that other deformations of
Plancherel measure have been used as a means to study Plancherel
measure itself -- see, e.g., \cite{johansson}. The phenomenon that a
small subtableau of a large random tableau has approximately the
Plancherel distribution was observed also in \cite{morsemckaywilf}
(see also \cite{stanley} for related results)
%, \cite{stanley}, \cite{regev}
for
% various other models of random tableaux.
a random tableau chosen uniformly among \emph{all} tableaux of size
$k$. Recently, it was shown \cite{pittel3} that the footprint of the
$k$ tallest stacks in a random unrestricted plane partition of high
volume also has in the limit the Plancherel distribution.
Another related observation is the easily checked fact that
$$ \sqrt{\alpha(1-\alpha)}\cdot
\tilde{g}_\alpha\left(\frac{u}{\sqrt{\alpha(1-\alpha)}}\right)
\xrightarrow[\alpha \searrow 0]{} \Omega(u),
$$
i.e. the shape of the level curves of our limit surface $L$ converges
after rescaling to the Plancherel limit curve $\Omega$, as one
approaches the corner of the square. This is consistent with Theorem
13, although is not formally implied by it, as here $k$ is a small
constant times $n^2$. It seems likely that in the regime when $k$
grows like $o(n^2)$, but much faster than $n^{2/3}$, $\nu_{n,k}$ and
$\mu_k$ become mutually singular, even though the limit shapes
coincide.
\subsection{The probability of a square plane partition to have all
parts distinct}
Denote by $M_{nN}$ the total number of $n\times n$ square plane
partitions of $N$. Let ${\cal M}_{nN}$ be the total number of those
partitions with all parts distinct. From Lemma 11 it follows that if
3312: \begin{equation}\label{eq:t1}
3313: \lim_{n,N\to\infty} \frac{n^6}{N}=0,
3314: \end{equation}
3315: then
3316: \begin{equation}\label{eq:t2}
3317: \lim_{n,N\to\infty}\frac{{\cal M}_{nN}}{M_{nN}}=1.
3318: \end{equation}
3319: Our goal is to show that \eqref{eq:t1} is essentially necessary for
3320: \eqref{eq:t2}. To motivate
3321: the statement, notice that
3322: the $k$-th largest part in a partition of $N$ into $n^2$ distinct parts is
3323: $n^2-k+1$, at least. So
3324: ${\cal M}_{nN}=0$ unless $N\ge n^2(n^2+1)/2$.
3325:
3326: \paragraph{Theorem 14.} Suppose that $n^4/N\to 0$. (i) If $\lim n^6/N=\infty$,
3327: then
3328: \begin{equation}\label{eq:t3}
3329: \lim_{n,N\to\infty}\frac{{\cal M}_{nN}}{M_{nN}}=0.
3330: \end{equation}
3331: (ii) If $\lim n^6/N=\alpha\in (0,\infty)$ then
3332: \begin{equation}\label{eq:t4}
3333: \lim_{n,N\to\infty}\frac{{\cal M}_{nN}}{M_{nN}}=e^{-\alpha/4}.
3334: \end{equation}
3335:
3336: \paragraph{Note.} Thus the reduction to the plane partitions with
3337: distinct parts used in the proof of Theorem 4 is valid if and only if
3338: $n^6=o(m)$.
3339:
3340: \paragraph{Proof sketch of Theorem 14.} We prove \eqref{eq:t3}, \eqref{eq:t4} by
3341: determining the asymptotic expressions of ${\cal M}_{nN}$ and $M_{nN}$.
3342:
3343: \paragraph{Part 1.}
3344: Begin with ${\cal M}_{nN}$. As in the proof of Lemma 10, we notice that
3345: -- given a linear partition of $N$ into $n^2$ distinct parts -- the
3346: number of the $n\times n$ square (descending) arrangements of these
3347: parts equals the total number of $n\times n$ square Young
3348: tableaux. So, denoting by $p_{nN}$ the total number of all such linear
3349: partitions, and by $d(\square_n)$ the number of all such tableaux, we
3350: obtain
3351: \begin{equation}\label{eq:t7}
3352: {\cal M}_{nN}=p_{nN}\cdot d(\square_n).
3353: \end{equation}
3354: Using the hook formula \eqref{eq:hook}, Euler's summation formula for $\sum_{s=1}^{n-1}
3355: (n-s)\log (n-s)$, and two identities for the Gamma
3356: function (see Bateman \cite{bateman}, Section 1.9)
3357: \begin{eqnarray*}
3358: \sum_{s=1}^ns\log s&=&\int_1^n\log \Gamma
3359: (x)\,dx+\frac{n(n+1)}{2}+\frac{n}{2}\log 2\pi,\\
3360: \log\Gamma (x)&=&
3361: \left(x-\frac{1}{2}\right)\log x-x+\frac{1}{2}\log 2\pi \\ & &
3362: +\int_0^{\infty}
3363: \left[(e^t-1)^{-1}-t^{-1}+\frac{1}{2}\right]t^{-1}e^{-tx}\,dt,\quad x>0.
3364: \end{eqnarray*}
3365: we obtain
3366: \begin{equation}\label{eq:t11}
3367: d(\square_n)\sim n^{11/12}\sqrt{2\pi}\exp\left(n^2\log n+n^2(-2\log
3368: 2+1/2)-\frac{1}{6}+\frac{\log 2}{12}
3369: -C\right),
3370: \end{equation}
3371: where
3372: \begin{equation}\label{eq:t10}
3373: C:=\int_0^{\infty}\left[(e^t-1)^{-1}-
3374: t^{-1}+\frac{1}{2}-\frac{t}{12}\right]t^{-2}e^{-t}\,dt.
3375: \end{equation}
3376: (A cruder formula
3377: $$
3378: d(\square_n)\sim \exp(n^2\log n+n^2(-2\log 2+1/2)+O(n\log n))
3379: $$
3380: was obtained, implicitly, in the proof of Lemma 1.)
3381:
3382: As for $p_{nN}$, the total number of partitions of $N$ into $n^2$ distinct
3383: parts, it is given by
3384: \begin{equation}\label{eq:t12}
3385: p_{nN}=[q^N t^{n^2}]\prod_{\ell=1}^{\infty}(1+q^{\ell}t).
3386: \end{equation}
3387: From a more general theorem of Vershik and Yakubovich \cite{vershikyak1},
3388: based on \eqref{eq:t12}, Fristedt's conditioning defice \cite{fristedt}, and an
3389: attendant local limit theorem result, it follows that
3390: \begin{equation}\label{eq:t13}
3391: p_{nN}\sim \frac{1}{2\pi N}
3392: \exp\left(n^2\log\frac{N}{n^4}+2n^2-
3393: \frac{n^6}{4N}(1+O(n^4/N))\right).
3394: \end{equation}
3395: Combining \eqref{eq:t11} and \eqref{eq:t13}, we arrive at
3396: \begin{multline}\label{eq:t14} \qquad\qquad
3397: {\cal M}_{nN}\sim\frac{n^{11/12}}{\sqrt{2\pi}N}\exp
\bigl(n^2\log(N/n^3)+n^2(-2\log 2+5/2)\\
-(n^6/4N)(1+O(n^4/N))+C^*\bigr), \qquad\qquad\\
C^*:=-\frac{1}{6}+\frac{\log 2}{12}-C, \qquad\qquad\qquad\qquad\qquad
\qquad\qquad\qquad\qquad\quad\ \
\end{multline}
with $C$ defined in \eqref{eq:t10}.
\paragraph{Part 2.} Turn now to $M_{nN}$, the number of all square $n\times n$
plane partitions of $N$. By the MacMahon formula for the number of
plane partitions with at most $n$ rows and $n$ columns,
\begin{equation}\label{eq:t15}
M_{nN}=[q^{N-n^2}]\prod_{\ell=1}^{\infty}(1-q^{\ell})^{-\ell}\cdot
\prod_{\ell>n}(1-q^{\ell})
^{2(\ell-n)}\cdot\prod_{\ell>2n}(1-q^{\ell})^{2n-\ell}.
%\tag t15
\end{equation}
(Alternatively, this formula follows from the hook expression for the
generating functions of plane partitions with a given shape discovered
by Stanley \cite{stanleyhook}.) We will use the techniques from
\cite{pittel3} inspired by Freiman's derivation of the main part of
Hardy-Ramanujan formula for the (linear) partition function, see
Postnikov \cite{postnikov}.
Let us take a close look at the generating function in \eqref{eq:t15}, which we
denote $p_n(q)$. Set $q=e^{-u}$, $\text{Re }u>0$. Taking logarithms,
using
\begin{equation}\label{eq:t16}
\log (1-e^{-mu})^{-1}=\sum_{j\ge 1}\frac{e^{-mju}}{j},
%\tag t16
\end{equation}
and changing the summation order, we obtain
\begin{multline}\label{eq:t17}
\qquad\quad\ \
\log p_n(e^{-u})=u\sum_{j\ge 1}\frac{1}{uj}\frac{e^{uj}}{(e^{uj}-1)^2}
(1-e^{-unj})^2\\ =
n^2\sum_{j=1}^{\infty}\frac{e^{-unj}}{j}
+u n^3\sum_{j=1}^{\infty}\frac{\psi(unj)}{(unj)^3}
\qquad\qquad\quad
\\
\qquad\qquad\quad\ \
-\frac{1}{12}\sum_{j=1}^{\infty}\frac{e^{-uj}}{j}(1-e^{-unj})^2
+u\sum_{j=1}^{\infty}\phi(uj)(1-e^{-unj})^2;\\
\phi(z):=\frac{e^z}{z(e^z-1)^2}-\frac{1}{z^3}+\frac{e^{-z}}{12z};
\qquad\qquad\qquad\qquad\quad\ \
3398: \\
3399: \psi(z):=(1-e^{-z})^2-z^2e^{-z}.
3400: \qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad
3401: %\endaligned\tag t17
3402: \end{multline}
3403: Using \eqref{eq:t16}, read backward, for the first and the third sums,
3404: and Euler's summation formula, with $m=1$, for the second and the
3405: fourth sum, we obtain from \eqref{eq:t17}: if $|u|^2n^4\to 0$, then
3406: \begin{multline}\label{eq:t19}
3407: \qquad\qquad
3408: p_n(e^{-u})\sim \frac{\exp\bigl(n^2(3/2-2\log 2)+(\log
3409: 2)/12+D\bigr)}{n^{1/12}}
3410: \cdot (1-e^{-nu})^{-n^2};\\
3411: D:=\int_0^{\infty}\phi(x)\,dx.
3412: \qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\ \
3413: %\endaligned\tag t19
3414: \end{multline}
3415: (According to Maple the integral of $\psi(x)$ equals $3/2-2\log 2$.)
3416: Now, by \eqref{eq:t15}, and the Cauchy formula
3417: $$
3418: M_{nN}=[q^{N-n^2}]p_n(q)=\frac{1}{2\pi i}
3419: \oint_{z=\rho e^{i\theta}\atop \theta\in (-\pi,\pi]}\frac
3420: {p_n(z)}{z^{N-n^2+1}}\,dz,
3421: $$
3422: where $\rho\in (0,1)$. In light of the last two equations, we set
3423: $\rho=e^{-u_0}$, and select $u_0$ that
3424: minimizes $-n^2\log (1-e^{-nu})+(N-n^2)u$, that is
3425: $$
3426: u_0=n^{-1}\log\left(1+\frac{n^3}{N-n^2}\right)\sim \frac{n^2}{N}.
3427: $$
Clearly $u_0^2n^4\to 0$.
Consider $|\theta|\le \theta_n=n^{-2}\epsilon_n$, $\epsilon_n=(n^4/N)^{1/2}$,
so that
$|\theta|^2n^4\to 0$ as well. It can be shown without much difficulty that
\begin{multline}\label{eq:t21}
\frac{1}{2\pi
i}\int\limits_{|\theta|\le\theta_n}\frac{dz}{(1-e^{-nu})^{n^2}z^{N-n^2+1}}\,dz
=\frac{1}
{2\pi
i}\int\limits_{u_0-i\theta_n}^{u_0+i\theta_n}\frac{e^{v(N-n^2)}}{(1-e^{-nv})^{n^2}}\,dv\\
=\frac{1}{n^{n^2}}\cdot\frac{N^{n^2-1}}{(n^2-1)!}(1+O(n^4/N))
\sim\frac{n}{\sqrt{2\pi}N}\left(\frac{eN}{n^3}\right)^{n^2}.
%\endaligned\tag t21
\end{multline}
(The last integral, extended to the closed contour obtained by connecting the
points $u_0\pm i\theta_n$
with a circular arc centered at the origin, is exactly
$$
\frac{1}{n^{n^2}}[t^{n^2-1}](1-nt)^{-(N-n^2+n+2)/n}=n^{-n^2}\binom{-n^{-1}(N-n^2+n+2)}{n^2-1}(-n)^{n^2-1},
$$
and the supplementary integral is less than this quantity by a factor
$(u_0/\theta_n)^{n^2}\sim
(n^4/N)^{n^2/2}$.)
So, using \eqref{eq:t19}, \eqref{eq:t21},
\begin{multline}\label{eq:t22}
\frac{1}{2\pi i}
\oint\limits_{z=\rho e^{i\theta}\atop \theta\in (-\theta_n,\theta_n]}\frac
{p_n(z)}{z^{N-n^2+1}}\,dz\\
\sim\frac{n^{11/12}}{\sqrt{2\pi}N}\exp\biggl(n^2\log\frac{N}{n^3}+n^2(5/2-2\log
2)
+(\log 2)/12+D+o(1)\biggr).
%\endmultline\tag t22
\end{multline}
The proof that the contribution of $\theta\notin [-\theta_n,\theta_n]$ is negligible
compared with the last expression is based on cruder estimates, not unlike those in
\cite{pittel3}, and we omit it. Therefore
\begin{equation}\label{eq:t26}
M_{nN}\sim\frac{n^{11/12}}{\sqrt{2\pi}N}\exp\bigg(n^2\log\frac{N}{n^3}+
n^2(5/2-2\log 2)+(\log 2)/12+D\bigg).
%\tag t26
\end{equation}
Comparing \eqref{eq:t26} and \eqref{eq:t14}, we see that
\begin{equation}\label{eq:t27}
\frac{{\cal M}_{nN}}{M_{nN}}=\exp\big(-(n^6/4N)(1+O(n^4/N))+A\big),
%\tag t27
\end{equation}
where, recalling the definition of $C$ and $D$,
\begin{eqnarray*}
A&=&-\frac{1}{6}-C-D\\
&=&-\frac{1}{6}-\int\limits_0^{\infty}\left[\left(\frac{1}{e^t-1}-\frac{1}{t}+\frac{1}{2}\right)t^{-2}e^{-t}
+\frac{e^t}{t(e^t-1)^2}-\frac{1}{t^3}\right]\,dt\\
&=&-\frac{1}{6}-\int\limits_0^{\infty}\frac{d}{dt}\left(-\frac{1}{t(e^t-1)}+\frac{1}{2t^2}+\frac{e^{-t}}
{2t^2}\right)\,dt\\
&=&-\frac{1}{6}
+\lim_{t\downarrow 0}\left(-\frac{1}{t(e^t-1)}+\frac{1}{2t^2}+
\frac{e^{-t}}{2t^2}\right)\\
&=&0.
\end{eqnarray*}
Thus we have
$$
\frac{{\cal M}_{nN}}{M_{nN}}=\exp\big(-(n^6/4N)(1+O(n^4/N))\big),
$$
which proves Theorem 14(i),(ii).
\qed
\subsection{Open problems}
We conclude with some open problems.
\begin{itemize}
\item {\bf Gaussian fluctuations.} Prove a central limit theorem for
the fluctuations of $g_{\lambda_T^{\lfloor \alpha n^2\rfloor}}$
around $\tilde{g}_\alpha$, and for the fluctuations of the
cotransition measure of $\lambda_T^{\lfloor \alpha n^2\rfloor}$ around
the semicircle distribution, in the spirit of \cite{ivanovolsh}.
\item {\bf Limiting distribution of $l_{n,k}(\pi)$.} Find a scaling
sequence $a_n$ and a distribution function $F$ such that, in the
notation of Theorem 3,
$$
\frac{l_{n,\lfloor \alpha n^2 \rfloor} - 2\sqrt{\alpha(1-\alpha)} n}{a_n}
\xrightarrow[n\to\infty]{\textrm{ in distribution }} F. $$
\item {\bf Limit surface for random Young tableaux of given shape.}
Prove a limit surface theorem for random Young tableaux of other
shapes. In general, one can consider any decreasing function
$f:[0,\infty)\to[0,\infty)$ such that $\int_0^\infty f(x)dx = 1$ as a
\emph{continual Young diagram}, i.e. as a limit of the rescaled graphs
of a sequence of Young diagrams of increasing sizes. We conjecture
that for each such continual diagram $f$, there should exist a limit
surface $L_f$, defined on the domain
$$ D_f := \{ (x,y) : x\ge 0, \ \,0 \le y \le f(x) \} $$
bounded between the $x$-axis and the graph of $f$, that describes the
asymptotic behavior of almost all random Young tableaux of shape
approximated by $f$.
\end{itemize}
\section*{Acknowledgements}
We thank Nati Linial for his suggestion that we work together, which
eventually led to the conception of this paper.
\begin{thebibliography}{02}
3428: \bibitem{aldousdiaconis} D. Aldous, P. Diaconis,
3429: \newblock Longest increasing subsequences: from patience sorting to
3430: the Baik-Deift-Johansson theorem.
3431: \newblock {\it Bull. Amer. Math. Soc.} 36 (1999), 413--432.
3432:
3433: \bibitem{apostol} T. M. Apostol,
3434: \newblock Introduction to Analytic Number Theory.
3435: \newblock Springer, 1976.
3436:
3437: \bibitem{baiketal} J. Baik, P. Deift, K. Johansson,
3438: \newblock On the distribution of the length of the longest increasing
3439: subsequence of random permutations.
3440: \newblock {\it J. Amer. Math. Soc.} 12 (1999), 1119--1178.
3441:
3442: \bibitem{baiketal2} J. Baik, P. Deift, K. Johansson,
3443: \newblock On the distribution of the length of the second row of a
3444: Young diagram under Plancherel measure.
3445: \newblock {\it Geom. Funct. Anal} 10 (2000), 1606--1607.
3446:
3447: \bibitem{bateman} H. Bateman, A. Erd\'elyi,
3448: \newblock Higher Transcendental Functions, vol. 1.
3449: \newblock McGraw-Hill, Ney York, 1953.
3450:
3451:
3452: \bibitem{borodinetal} A. Borodin, A. Okounkov, G. Olshanski,
3453: \newblock Asymptotics of Plancherel measures for symmetric groups.
3454: \newblock {\it J. Amer. Math. Soc.} 13 (2000), 481--515.
3455:
3456: \bibitem{cerfkenyon} R. Cerf, R. Kenyon,
3457: \newblock The low-temperature expansion of the Wulff crystal in the 3D
3458: Ising model.
3459: \newblock {\it Comm. Math. Phys} 222 (2001), 147--179.
3460:
3461: \bibitem{cohnlarspropp} H. Cohn, M. Larsen, J. Propp,
3462: \newblock The shape of a typical boxed plane partition.
3463: \newblock {\it New York J. Math.} 4 (1998), 137--165.
3464:
3465: \bibitem{erdoslehner} P. Erd\"os, J. Lehner,
3466: \newblock The distribution of the number of summands in the partitions
3467: of a positive integer.
3468: \newblock {\it Duke J. Math.} 8 (1941), 335--345.
3469:
3470: \bibitem{estradakanwal} R. Estrada, R. P. Kanwal,
\newblock Singular Integral Equations.
\newblock Birkh\"auser, 2000.
\bibitem{framehook} J. S. Frame, G. de B. Robinson, R. M. Thrall,
\newblock The hook graphs of the symmetric groups.
\newblock {\it Canadian J. Math.} 6 (1954), 316--324.
\bibitem{fristedt} B. Fristedt,
\newblock The structure of random partitions of large integers,
\newblock {\it Trans. Amer. Math. Soc.} 337 (1993), 703--735.
\bibitem{greenenij} C. Greene, A. Nijenhuis, H. Wilf,
\newblock A probabilistic proof of a formula for the number of Young
tableaux of a given shape.
\newblock {\it Adv. Math.} 31 (1979), 104--109.
\bibitem{gribov} A. B. Gribov,
\newblock The limit Young tableau with respect to the Plancherel
measure.
\newblock (Russian) Vestnik
Leningrad. Univ. Mat. Mekh. Astronom. 1986, vyp. 2, 100--102, 131.
\bibitem{hardyetal} G. H. Hardy, J. E. Littlewood, G. P\'olya,
\newblock Inequalities, 2nd. ed.
\newblock Cambridge University Press, Cambridge, 1952.
\bibitem{ivanovolsh} V. Ivanov, G. Olshanski,
\newblock Kerov's central limit theorem for the Plancherel measure on
Young diagrams.
\newblock In: Symmetric Functions 2001: Surveys of Developments and
Perspectives. Proc. NATO Advanced Study Institute, ed. S. Fomin,
Kluwer, 2002.
\bibitem{johansson} K. Johansson,
\newblock Discrete orthogonal polynomial ensembles and the Plancherel
measure.
\newblock {\it Ann. of Math.} 153 (2001), 259--296.
%\bibitem{kanwal} R. P. Kanwal,
%\newblock Linear Integral Equations.
%\newblock Academic Press, New York, 1971.
%\bibitem{kenyonokounkovsheff} R. Kenyon, A. Okounkov, S. Sheffield,
%\newblock
3471:
3472: \bibitem{kerov} S. V. Kerov,
3473: \newblock A differential model of growth of Young diagrams.
3474: \newblock (Russian) Proceedings of the St. Petersburg Mathematical
3475: Society, vol. IV, 111--130; translation in {\it
3476: Amer. Math. Soc. Transl. Ser. 2}, 188, Amer. Math. Soc., Providence, RI.
3477:
3478: \bibitem{knuth}
3479: D. E. Knuth,
3480: \newblock The Art of Computer Programming, vol. 3: Sorting and
3481: Searching, 2nd. ed.
3482: \newblock Addison-Wesley, 1998.
3483:
3484: \bibitem{loganshepp} B. F. Logan, L. A. Shepp,
3485: \newblock A variational problem for random Young tableaux.
3486: \newblock {\it Adv. Math.} 26 (1977), 206--222.
3487:
3488: \bibitem{lulovpittel} N. Lulov, B. Pittel,
3489: \newblock On the random Young diagrams and their cores.
3490: \newblock {\it J. Combin. Theory Ser. A.} 86 (1999), 245--280.
3491:
3492: \bibitem{morsemckaywilf} B. D. McKay, J. Morse, H. Wilf,
3493: \newblock The distributions of the entries of Young tableaux.
3494: \newblock {\it J. Combin. Theory Ser. A} 97 (2002), 117--128.
3495:
3496: \bibitem{oconnell} N. O'Connell,
3497: \newblock A path-transformation for random walks and the
3498: Robinson-Schensted correspondence.
3499: \newblock {\it Trans. Amer. Math. Soc.} 355 (2003), 3669--3697.
3500:
3501: \bibitem{okounkov} A. Okounkov,
3502: \newblock Random matrices and random permutations.
3503: \newblock {\it Internat. Math. Res. Notices} 2000:20 (2000), 1043--1095.
3504:
3505: \bibitem{pittel1} B. Pittel,
3506: \newblock On a likely shape of the random Ferrers diagram.
3507: \newblock {\it Adv. Appl. Math.} 18 (1997), 432--488.
3508:
3509: \bibitem{pittel3} B. Pittel,
3510: \newblock On dimensions of a random solid diagram.
3511: \newblock preprint.
3512:
3513: \bibitem{pittel2} B. Pittel,
3514: \newblock On the distribution of the number of Young tableaux for a
3515: uniformly random diagram.
3516: \newblock {\it Adv. Appl. Math.} 29 (2002), 184--214.
3517:
3518: \bibitem{porterstirling} D. Porter, D. S. G. Stirling,
3519: \newblock Integral Equations.
3520: \newblock Cambridge University Press, Cambridge, 1990.
3521:
3522: \bibitem{postnikov} A. G. Postnikov,
3523: \newblock Introduction to Analytic Number Theory,
3524: \newblock Translations of Mathematical Monographs 68,
3525: \newblock American Mathematical Society, Providence, RI, 1988.
3526:
3527: %\bibitem{regev} A. Regev,
3528: %\newblock
3529:
3530: \bibitem{romik} D. Romik,
3531: \newblock Explicit formulas for hook walks on continual Young
3532: diagrams.
3533: \newblock {\it Adv. Appl. Math.} 32 (2004), 625--654.
3534:
3535: \bibitem{schensted} C. Schensted,
3536: \newblock Longest increasing and decreasing subsequences.
3537: \newblock {\it Canadian J. Math.} 13 (1961), 179--191.
3538:
3539: \bibitem{stanley} R. P. Stanley,
3540: \newblock On the enumeration of skew Young tableaux.
3541: \newblock {\it Adv. Appl. Math.} 30 (2003), 283--294.
3542:
3543: \bibitem{stanleyhook} R. P. Stanley,
3544: \newblock Theory and application of plane partitions, II.
3545: \newblock {\it Studies in Appl. Math} 50 (1971), 259--279.
3546:
3547: \bibitem{titchmarsh} E. C. Titchmarsh,
3548: \newblock Introduction to the Theory of Fourier Integrals.
3549: \newblock Oxford University Press, Oxford, 1937.
3550:
3551: \bibitem{vershikkerov1} A. M. Vershik, S. V. Kerov,
3552: \newblock Asymptotics of the Plancherel measure of the symmetric group
3553: and the limiting shape of Young tableaux.
3554: \newblock {\it Soviet Math. Dokl.} 18 (1977), 527--531.
3555:
3556: \bibitem{vershikkerov2} A. M. Vershik, S. V. Kerov,
3557: \newblock The asymptotics of maximal and typical dimensions of
3558: irreducible representations of the symmetric group.
3559: \newblock {\it Funct. Anal. Appl.} 19 (1985), 21--31.
3560:
3561: \bibitem{vershikyak1} A. M. Vershik, Yu. Yakubovich,
3562: \newblock The limit shape and fluctuations of random partitions of
3563: naturals with fixed number of summands.
3564: \newblock {\it Mosc. Math. J.} 1 (2001), 457--468, 472.
3565:
3566: \bibitem{vershikyak2} A. M. Vershik, Yu. Yakubovich,
3567: \newblock Asymptotics of the uniform measures on simplices and random
3568: compositions and partitions.
3569: \newblock {\it Funct. Anal. Appl.} 37 (2003), 273--280.
3570:
3571: \end{thebibliography}
3572:
3573: \end{document}
3574: