This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Breaking Multivariate Records

James Allen Fill Department of Applied Mathematics and Statistics, The Johns Hopkins University, 3400 N. Charles Street, Baltimore, MD 21218-2682 USA [email protected] http://www.ams.jhu.edu/\simfill/
(Date: September 29, 2021; revised March 7, 2023)
Abstract.

For a sequence of i.i.d. dd-dimensional random vectors with independent continuously distributed coordinates, say that the nnth observation in the sequence sets a record if it is not dominated in every coordinate by an earlier observation; for jnj\leq n, say that the jjth observation is a current record at time nn if it has not been dominated in every coordinate by any of the first nn observations; and say that the nnth observation breaks kk records if it sets a record and there are kk observations that are current records at time n1n-1 but not at time nn.

For general dimension dd, we identify, with proof, the asymptotic conditional distribution of the number of records broken by an observation given that the observation sets a record.

Fix dd, and let 𝒦(d)\mathcal{K}(d) be a random variable with this distribution. We show that the (right) tail of 𝒦(d)\mathcal{K}(d) satisfies

(𝒦(d)k)exp[Ω(k(d1)/(d2+d3))] as k\operatorname{\mathbb{P}{}}(\mathcal{K}(d)\geq k)\leq\exp\left[-\Omega\!\left(k^{(d-1)/(d^{2}+d-3)}\right)\right]\mbox{\ \ as $k\to\infty$}

and

(𝒦(d)k)exp[O(k1/(d1))] as k.\operatorname{\mathbb{P}{}}(\mathcal{K}(d)\geq k)\geq\exp\left[-O\!\left(k^{1/(d-1)}\right)\right]\mbox{\ \ as $k\to\infty$}.

When d=2d=2, the description of 𝒦(2)\mathcal{K}(2) in terms of a Poisson process agrees with the main result from [3] that 𝒦(2)\mathcal{K}(2) has the same distribution as 𝒢1\mathcal{G}-1, where 𝒢Geometric(1/2)\mathcal{G}\sim\mbox{\rm Geometric$(1/2)$}. Note that the lower bound on (𝒦(d)k)\operatorname{\mathbb{P}{}}(\mathcal{K}(d)\geq k) implies that the distribution of 𝒦(d)\mathcal{K}(d) is not (shifted) Geometric for any d3d\geq 3.

We show that (𝒦(d)1)=exp[Θ(d)]\operatorname{\mathbb{P}{}}(\mathcal{K}(d)\geq 1)=\exp[-\Theta(d)] as dd\to\infty; in particular, 𝒦(d)0\mathcal{K}(d)\to 0 in probability as dd\to\infty.

Key words and phrases:
Multivariate records, Pareto records, record breaking, current records, maxima, Poisson point processes, asymptotics
2010 Mathematics Subject Classification:
Primary: 60D05; Secondary: 60F05
Research supported by the Acheson J. Duncan Fund for the Advancement of Research in Statistics.

1. Introduction and main result

Let 𝐗(1),𝐗(2),\mathbf{X}^{(1)},\mathbf{X}^{(2)},\dots be i.i.d. (independent and identically distributed) copies of a random vector 𝐗=(X1,,Xd)\mathbf{X}=(X_{1},\ldots,X_{d}) with independent coordinates, each uniformly distributed over the unit interval. In this paper, for general dimension dd, we prove (Theorem 1.3) that the conditional distribution of the number of records broken by the nnth observation 𝐗(n)\mathbf{X}^{(n)} given that 𝐗(n)\mathbf{X}^{(n)} sets a record has a weak limit [call it μ(d)\mu(d), the distribution or law (𝒦(d)){\mathcal{L}}(\mathcal{K}(d)) of a random variable 𝒦(d)\mathcal{K}(d)] as nn\to\infty, and we identify this limit. (See Definition 1.1 for the relevant definitions regarding records.) The case d=1d=1 is trivial: 𝒦(d)=1\mathcal{K}(d)=1 with probability 11. The case d=2d=2 is treated in depth in [3], where it is shown that μ(2)=(𝒢1)\mu(2)={\mathcal{L}}(\mathcal{G}-1) with 𝒢Geometric(1/2)\mathcal{G}\sim\mbox{\rm Geometric$(1/2)$}.

Notation. We let 𝟏(E)=1 or 0{\bf 1}(E)=\mbox{$1$ or $0$} according as EE is true or false, and for a random variable YY and an event AA we write 𝔼(Y;A)\operatorname{\mathbb{E}{}}(Y;A) as shorthand for 𝔼[Y𝟏(A)]\operatorname{\mathbb{E}{}}[Y{\bf 1}(A)]. We write ln\ln or L\operatorname{L} for natural logarithm. For dd-dimensional vectors 𝐱=(x1,,xd)\mathbf{x}=(x_{1},\dots,x_{d}) and 𝐲=(y1,,yd)\mathbf{y}=(y_{1},\dots,y_{d}), write 𝐱𝐲\mathbf{x}\prec\mathbf{y} to mean that xj<yjx_{j}<y_{j} for j=1,,dj=1,\dots,d. The notation 𝐱𝐲\mathbf{x}\succ\mathbf{y} means 𝐲𝐱\mathbf{y}\prec\mathbf{x}. The notations 𝐱𝐲\mathbf{x}\preceq\mathbf{y} and 𝐲𝐱\mathbf{y}\succeq\mathbf{x} each mean that either 𝐱𝐲\mathbf{x}\prec\mathbf{y} or 𝐱=𝐲\mathbf{x}=\mathbf{y}. Whenever we speak of incomparable vectors, we will mean vectors belonging to d\mathbb{R}^{d} that are (pairwise-)incomparable with respect to the partial order \preceq. We write x+:=j=1dxjx_{+}:=\sum_{j=1}^{d}x_{j} for the sum of coordinates of x=(x1,,xd)x=(x_{1},\dots,x_{d}) and \|\cdot\| for 1\ell^{1}-norm; thus x+=𝐱x_{+}=\|\mathbf{x}\| if 𝐱𝟎\mathbf{x}\succ{\bf 0}, where 𝟎{\bf 0} is the vector (0,,0)(0,\dots,0).

We begin with some relevant definitions, taken from [5; 4; 3]. We give definitions for record-large values; we consider such records exclusively except in the case of Theorem 1.3, which deals with record-small values (for which the analogous definitions are obvious).

Let 𝐗(1),𝐗(2),\mathbf{X}^{(1)},\mathbf{X}^{(2)},\dots be i.i.d. (independent and identically distributed) copies of a random vector 𝐗\mathbf{X} with independent coordinates, each drawn from a common specified distribution. We mainly consider the case (as in [5; 4]) where the common distribution is Exponential(1)(1), but in Theorem 1.3 we consider the uniform distribution over the unit interval. As far as counting records or broken records, any continuous distribution gives the same results.

Definition 1.1.

(a) We say that 𝐗(n)\mathbf{X}^{(n)} is a Pareto record (or simply record, or that 𝐗(n)\mathbf{X}^{(n)} sets a record at time nn) if 𝐗(n)𝐗(i)\mathbf{X}^{(n)}\not\prec\mathbf{X}^{(i)} for all 1i<n1\leq i<n.

(b) If 1jn1\leq j\leq n, we say that 𝐗(j)\mathbf{X}^{(j)} is a current record (or remaining record, or maximum) at time nn if 𝐗(j)𝐗(i)\mathbf{X}^{(j)}\not\prec\mathbf{X}^{(i)} for all i[n]i\in[n].

(c) If 0kn0\leq k\leq n, we say that 𝐗(n)\mathbf{X}^{(n)} breaks (or kills) kk records if 𝐗(n)\mathbf{X}^{(n)} sets a record and there exist precisely kk values jj with 1j<n1\leq j<n such that 𝐗(j)\mathbf{X}^{(j)} is a current record at time n1n-1 but is not a current record at time nn.

(d) For n1n\geq 1 (or n0n\geq 0, with the obvious conventions) let RnR_{n} denote the number of records 𝐗(k)\mathbf{X}^{(k)} with 1kn1\leq k\leq n, and let rnr_{n} denote the number of remaining records at time nn.

Remark 1.2.

Definition 1.1 is illustrated in Figures 12, adapted from [3, Figure 1] and [4, Figure 2], respectively. In dimension 22, remaining records can be ordered northwest to southeast, as seen in Figure 1.

​​​​0

Figure 1. In this 22-dimensional example, after n1n-1 observations, none of which fall in the shaded region, there are rn1=6r_{n-1}=6 remaining records. The nthn^{\rm\scriptsize th} observation, shown in bright green at the intersection of the dashed lines, breaks the Kn=k=3K_{n}=k=3 remaining records shown in red but not the rn1Kn=3r_{n-1}-K_{n}=3 remaining records shown in dark green.

In dimensions d3d\geq 3, the record-setting region is a (typically) more complicated set, as illustrated in Figure 2.

​​​​0x1x_{1}x2x_{2}x3x_{3}gg

Figure 2. In this 33-dimensional example, suppose that there are 88 remaining records, as shown in green. A new observation will set a record if and only if it belongs to the union of positive orthants translated by one of the 1717 points shown in purple. The lower boundary of one of these translated orthants (corresponding to the purple point labeled gg) is shown using dashed lines.

A maximum 𝐱\mathbf{x} for a given set SdS\subseteq\mathbb{R}^{d} is defined in similar fashion to Definition 1.1(b): We say that 𝐱\mathbf{x} is a maximum of SS if 𝐱𝐲\mathbf{x}\not\prec\mathbf{y} for all 𝐲S\mathbf{y}\in S. In this language, if 1jn1\leq j\leq n we have that 𝐗(j)\mathbf{X}^{(j)} is a remaining record at time nn if 𝐗(j)\mathbf{X}^{(j)} is a maximum of {𝐗(1),𝐗(2),,𝐗(n)}\{\mathbf{X}^{(1)},\mathbf{X}^{(2)},\ldots,\mathbf{X}^{(n)}\}.

Here is the main result of this paper, illustrated in Figure 3.

​​​​0𝐱\mathbf{x}𝐙(1)\mathbf{Z}^{(1)}𝐙(2)\mathbf{Z}^{(2)}

Figure 3. In this (d=2d=2) realization of the Poisson point process in Theorem 1.3 and the theorem’s possible extension in Section 2, the maxima are shown in red [and labeled 𝐙(1)\mathbf{Z}^{(1)} and 𝐙(2)\mathbf{Z}^{(2)}] if dominated by the bright green point 𝐱\mathbf{x} (so 𝒦=2\mathcal{K}=2) and in dark green otherwise. The intensity of the process vanishes in the shaded region.
Theorem 1.3.

Let 𝐗(1),𝐗(2),\mathbf{X}^{(1)},\mathbf{X}^{(2)},\dots be i.i.d. dd-variate observations, each with independent Exponential(1)(1) coordinates. Let Kn=1K_{n}=-1 if 𝐗(n)\mathbf{X}^{(n)} does not set a record, and otherwise let KnK_{n} denote the number of remaining records killed by 𝐗(n)\mathbf{X}^{(n)}. Then KnK_{n}, conditionally given Kn0K_{n}\geq 0, converges in distribution as nn\to\infty to the law of 𝒦𝒦(d)\mathcal{K}\equiv\mathcal{K}(d), where

(𝒦) is the mixture 𝔼(𝒦G).\mbox{\rm${\mathcal{L}}(\mathcal{K})$ is the mixture $\operatorname{\mathbb{E}{}}{\mathcal{L}}(\mathcal{K}_{G})$}.

Here GG is distributed standard Gumbel and

𝒦g=number of maxima dominated by 𝐱𝐱(g):=(g/d,,g/d)\mathcal{K}_{g}=\mbox{\rm number of maxima dominated by $\mathbf{x}\equiv\mathbf{x}(g):=(g/d,\ldots,g/d)$}

in a nonhomogeneous Poisson point process in d\mathbb{R}^{d} with intensity function

𝟏(𝐳𝐱)ez+,𝐳d.{\bf 1}(\mathbf{z}\not\succ\mathbf{x})e^{-z_{+}},\quad\mathbf{z}\in\mathbb{R}^{d}.

It is easy to see from the form of the intensity function that, for each gg\in\mathbb{R}, the distribution of 𝒦g\mathcal{K}_{g} is unchanged if 𝐱(g)\mathbf{x}(g) is replaced by any point 𝐱~(g)d\tilde{\mathbf{x}}(g)\in\mathbb{R}^{d} with x~(g)+=g\tilde{x}(g)_{+}=g.

By using the decreasing bijection zezz\mapsto e^{-z} from \mathbb{R} to +:=(0,)\mathbb{R}_{+}:=(0,\infty), which is also a bijection from +\mathbb{R}_{+} to (0,1)(0,1), Theorem 1.3 can be recast equivalently as follows:

Theorem 1.3. Suppose that independent dd-variate observations, each uniformly distributed in the unit hypercube (0,1)d(0,1)^{d}, arrive at times 1,2,1,2,\ldots, and consider record-small values. Let Kn=1K_{n}=-1 if the nthn^{\rm\scriptsize th} observation is not a new record, and otherwise let KnK_{n} denote the number of remaining records killed by the nthn^{\rm\scriptsize th} observation. Then KnK_{n}, conditionally given Kn0K_{n}\geq 0, converges in distribution as nn\to\infty to the law of 𝒦𝒦(d)\mathcal{K}\equiv\mathcal{K}(d), where

(𝒦) is the mixture 𝔼(𝒦W).\mbox{\rm${\mathcal{L}}(\mathcal{K})$ is the mixture $\operatorname{\mathbb{E}{}}{\mathcal{L}}(\mathcal{K}^{\prime}_{W})$}.

Here WW is distributed Exponential(1)(1) and

𝒦w=number of minima that dominate 𝐱𝐱(w):=(w1/d,,w1/d)\mathcal{K}^{\prime}_{w}=\mbox{\rm number of minima that dominate $\mathbf{x}^{\prime}\equiv\mathbf{x}^{\prime}(w):=(w^{1/d},\ldots,w^{1/d})$}

in a (homogeneous) unit-rate Poisson point process in the set

{𝐳+d:𝐳𝐱}.\{\mathbf{z}\in\mathbb{R}^{d}_{+}:\mathbf{z}\not\prec\mathbf{x}^{\prime}\}.

Section 2 gives a non-rigorous but informative proof of Theorem 1.3, and the much longer Section 3 gives a rigorous proof. Tail probabilities for 𝒦(d)\mathcal{K}(d) are studied (asymptotically) for each fixed dd in Section 4; moments are considered, too. Asymptotics of (𝒦(d)){\mathcal{L}}(\mathcal{K}(d)) as dd\to\infty are studied in Section 5; in particular, we find that 𝒦(d)\mathcal{K}(d) converges in probability to 0. Finally, in Section 6 we show that the main Theorem 1.3 of this paper reduces to the main Theorem 1.1 of [4] when d=2d=2.

2. Informal proof of main theorem

In this section we give an informal (heuristic) proof of Theorem 1.3; a formal proof is given in the next section. The informal proof suggests that it should be possible to prove extensions of Theorem 1.3 (under the same hypotheses as the theorem and using the same notation) such as the following, but we haven’t pursued such extensions in this paper.

Possible extension. Let k1k\geq 1. Over the event Kn=kK_{n}=k, let 𝐘(n,1),\mathbf{Y}^{(n,1)},\ldots, 𝐘(n,k)\mathbf{Y}^{(n,k)} denote the values of the kk records killed by 𝐗(n)\mathbf{X}^{(n)}, listed (for definiteness) in increasing order of first coordinate, and let 𝐘(n,):=𝐗(n)\mathbf{Y}^{(n,\ell)}:=\mathbf{X}^{(n)} for >k\ell>k. Then (𝐗(n)𝐘(n,1),𝐗(n)𝐘(n,2),)(\mathbf{X}^{(n)}-\mathbf{Y}^{(n,1)},\mathbf{X}^{(n)}-\mathbf{Y}^{(n,2)},\dots) converges in distribution to

(𝐱(G)𝐙(1)(G),𝐱(G)𝐙(2)(G),);(\mathbf{x}(G)-\mathbf{Z}^{(1)}(G),\mathbf{x}(G)-\mathbf{Z}^{(2)}(G),\dots);

here, over the event {G=g,𝒦=k}\{G=g,\,\mathcal{K}=k\}, the points 𝐙(1)(g)\mathbf{Z}^{(1)}(g), ,𝐙(k)(g)\ldots,\mathbf{Z}^{(k)}(g) are the kk maxima counted by 𝒦g\mathcal{K}_{g}, listed in increasing order of first coordinate, and 𝐙()(g):=𝐱(g)\mathbf{Z}^{(\ell)}(g):=\mathbf{x}(g) for >k\ell>k.

Before beginning our heuristic proof of Theorem 1.3, we gather and utilize important information from [5] in the following remark.

Remark 2.1.

(a) It is well known that

pn:=(𝐗(n)sets a record)=(Kn0)1n(lnn)d1(d1)!p_{n}:=\operatorname{\mathbb{P}{}}(\mathbf{X}^{(n)}\leavevmode\nobreak\ \text{sets a record})=\operatorname{\mathbb{P}{}}(K_{n}\geq 0)\sim\frac{1}{n}\frac{(\ln n)^{d-1}}{(d-1)!} (2.1)

as nn\to\infty; see, for example, [5, (4.5)].

(b) Recall from [5, proof of Theorem 1.4] that, conditionally given Kn0K_{n}\geq 0, the joint density of

Gn=𝐗(n)Ln,𝐔n=𝐗(n)1𝐗(n)G_{n}=\|\mathbf{X}^{(n)}\|-\operatorname{L}n,\quad\mathbf{U}_{n}=\|\mathbf{X}^{(n)}\|^{-1}\mathbf{X}^{(n)}

with respect to the product of Lebesgue measure on \mathbb{R} and uniform distribution on the probability simplex 𝒮d1{\mathcal{S}}_{d-1} is

(g,𝐮)pn1n1(g+Ln)d1(d1)!eg(1n1eg)n1𝟏(g>Ln)(g,\mathbf{u})\mapsto p_{n}^{-1}n^{-1}\frac{(g+\operatorname{L}n)^{d-1}}{(d-1)!}e^{-g}(1-n^{-1}e^{-g})^{n-1}{\bf 1}(g>-\operatorname{L}n) (2.2)

and, as nn\to\infty, converges pointwise to

(g,𝐮)egexp(eg),(g,\mathbf{u})\mapsto e^{-g}\exp\left(-e^{-g}\right),

the density (with respect to the same product measure) of the standard Gumbel probability measure and uniform measure on 𝒮d1{\mathcal{S}}_{d-1}. Thus by Scheffé’s theorem (e.g., [1, Theorem 16.12]), there is total variation convergence of (Gn,𝐔n){\mathcal{L}}(G_{n},\mathbf{U}_{n}) to (G,𝐔){\mathcal{L}}(G,\mathbf{U}) in obvious notation.

(c) We also claim that

𝔼eGne2gexp(eg)dg=1\operatorname{\mathbb{E}{}}e^{-G_{n}}\to\int_{-\infty}^{\infty}e^{-2g}\exp(-e^{-g})\,\mathrm{d}g=1 (2.3)

as nn\to\infty. To see this, first observe that as nn\to\infty we have [using (2.2) and (2.1)] that

𝔼eGn\displaystyle\operatorname{\mathbb{E}{}}e^{-G_{n}} =𝟏(g>Ln)pn1(g+Ln)d1(d1)!e2gn(1n1eg)n1dg\displaystyle=\int_{-\infty}^{\infty}\!{\bf 1}(g>-\operatorname{L}n)\,p_{n}^{-1}\,\frac{(g+\operatorname{L}n)^{d-1}}{(d-1)!}\,\frac{e^{-2g}}{n}\,\left(1-n^{-1}e^{-g}\right)^{n-1}\,\mathrm{d}g
𝟏(g>Ln)(1+gLn)d1e2g(1n1eg)n1dg.\displaystyle\sim\int_{-\infty}^{\infty}\!{\bf 1}(g>-\operatorname{L}n)\,\left(1+\frac{g}{\operatorname{L}n}\right)^{d-1}\,e^{-2g}\,\left(1-n^{-1}e^{-g}\right)^{n-1}\,\mathrm{d}g.

For fixed gg\in\mathbb{R}, as nn\to\infty the integrand converges to

e2gexp(eg),e^{-2g}\exp\left(-e^{-g}\right),

so it suffices to invoke the dominated convergence theorem. For n3n\geq 3 we can dominate the integrand by

(1+|g|)d1e2zexp(n1neg)(1+|g|)d1e2gexp(23eg),(1+|g|)^{d-1}e^{-2z}\exp\left(-\frac{n-1}{n}e^{-g}\right)\leq(1+|g|)^{d-1}e^{-2g}\exp\left(-\frac{2}{3}e^{-g}\right),

and this last expression integrates: Indeed, as gg\to-\infty, it is asymptotically equivalent to |g|d1e2|g|exp(23e|g|)|g|^{d-1}e^{2|g|}\exp\left(-\frac{2}{3}e^{|g|}\right), and as gg\to\infty, it is asymptotically equivalent to gd1e2gg^{d-1}e^{-2g}.

Here now is a heuristic proof of Theorem 1.3. Recall that we use the shorthand notation L\operatorname{L} for natural logarithm. From Remark 2.1(a), the conditional density of Gn=X+(n)LnG_{n}=X^{(n)}_{+}-\operatorname{L}n given Kn0K_{n}\geq 0 converges pointwise to the standard Gumbel density as nn\to\infty; further, the conditional distribution of 𝐗(n)\mathbf{X}^{(n)} given Kn0K_{n}\geq 0 and Gn=gG_{n}=g is uniform over all positive dd-tuples summing to Ln+g\operatorname{L}n+g. Next, given Kn0K_{n}\geq 0 and 𝐗(n)=𝐰0\mathbf{X}^{(n)}=\mathbf{w}\succ 0 with w+=Ln+gw_{+}=\operatorname{L}n+g (call this condition CC), the random variables 𝐗(1)𝐰,,𝐗(n1)𝐰\mathbf{X}^{(1)}-\mathbf{w},\ldots,\mathbf{X}^{(n-1)}-\mathbf{w} are i.i.d., each with (conditional) density proportional to 𝐳ez+\mathbf{z}\mapsto e^{-z_{+}} with respect to Lebesgue measure on

Sn:={𝐳d:𝐳𝐰 and 𝐳𝟎}={𝐳d:𝐳𝐰}+d,S_{n}:=\{\mathbf{z}\in\mathbb{R}^{d}:\mathbf{z}\succ-\mathbf{w}\mbox{\ and\ }\mathbf{z}\not\succ{\bf 0}\}=\{\mathbf{z}\in\mathbb{R}^{d}:\mathbf{z}\succ-\mathbf{w}\}-\mathbb{R}_{+}^{d},

a proper set difference; the proportionality constant is then the reciprocal of ew+1=neg1e^{w_{+}}-1=ne^{g}-1. For Δ>0\Delta>0 and gg fixed real numbers, this implies that the conditional density in question evaluated at a point 𝐳0\mathbf{z}\not\succ 0 at distance (say, 1\ell_{1}-distance) at most Δ\Delta from 𝟎{\bf 0} is asymptotically equivalent to n1ez+gn^{-1}e^{-z_{+}-g}.

It follows that (still conditionally given CC) the set of points 𝐗(i)𝐰+𝐱(g)\mathbf{X}^{(i)}-\mathbf{w}+\mathbf{x}(g) with i{1,,n1}i\in\{1,\ldots,n-1\} that are within distance Δ\Delta of 𝐱(g)\mathbf{x}(g) is approximately distributed (when nn is large) as a nonhomogeneous Poisson point process with intensity function as described in Theorem 1.3, restricted to the ball of radius Δ\Delta centered at 𝐱(g)\mathbf{x}(g). The maxima of this restricted Poisson process ought to be the same as the maxima of the unrestricted process when Δ\Delta is sufficiently large. Letting Δ\Delta\to\infty, Theorem 1.3 results.

3. Proof of Theorem 1.3

In this section we give a complete formal proof of the main Theorem 1.3. Let μn:=(KnKn0)\mu_{n}:={\mathcal{L}}(K_{n}\mid K_{n}\geq 0). In Subsection 3.1 we prove the existence of a weak limit μ\mu for μn\mu_{n}, and in Subsection 3.2 we identify μ\mu as the law of 𝒦\mathcal{K} as described in Theorem 1.3.

3.1. Existence of weak limit

In this subsection we prove the following proposition.

Proposition 3.1.

There exists a probability measure μ\mu on {1,2,}\{1,2,\dots\} such that μn=(KnKn0)\mu_{n}={\mathcal{L}}(K_{n}\mid K_{n}\geq 0) converges weakly to μ\mu.

Indeed, Proposition 3.1 is established by showing that the sequence (μn)(\mu_{n}) is tight (Lemma 3.2) and has a vague limit (Lemma 3.3).

Lemma 3.2.

(i) We have 𝔼(KnKn0)1\operatorname{\mathbb{E}{}}(K_{n}\mid K_{n}\geq 0)\to 1 as nn\to\infty.

(ii) The sequence (μn)(\mu_{n}) is tight.

Proof.

By Markov’s inequality, boundedness of first moments is a sufficient condition for tightness. Thus (ii) follows from (i).

We now prove (i). When d=1d=1 we have Kn=1K_{n}=1 deterministically if Kn0K_{n}\geq 0, so (i) is obvious.

We therefore assume henceforth that d2d\geq 2. Recalling Definition 1.1(d) and introducing the dimension dd into the notation, denote the number of records that have been broken through time nn by βn(d):=Rn(d)rn(d)\beta_{n}(d):=R_{n}(d)-r_{n}(d). Then

𝔼(Kn|Kn0)\displaystyle\operatorname{\mathbb{E}{}}(K_{n}\,|\,K_{n}\geq 0) =𝔼[βn(d)βn1(d)|𝐗(n)sets a record]\displaystyle=\operatorname{\mathbb{E}{}}[\beta_{n}(d)-\beta_{n-1}(d)\,|\,\mathbf{X}^{(n)}\leavevmode\nobreak\ \text{sets a record}]
=𝔼[βn(d)βn1(d);𝐗(n)sets a record](𝐗(n)sets a record)\displaystyle=\dfrac{\operatorname{\mathbb{E}{}}[\beta_{n}(d)-\beta_{n-1}(d);\,\mathbf{X}^{(n)}\leavevmode\nobreak\ \text{sets a record}]}{\operatorname{\mathbb{P}{}}(\mathbf{X}^{(n)}\leavevmode\nobreak\ \text{sets a record})}
=𝔼[βn(d)βn1(d)]d(𝐗(n)sets a record).\displaystyle=\frac{\operatorname{\mathbb{E}{}}[\beta_{n}(d)-\beta_{n-1}(d)]}{\operatorname{\mathbb{P}{}}_{d}(\mathbf{X}^{(n)}\leavevmode\nobreak\ \text{sets a record})}. (3.1)

For the numerator of (3.1) we observe

𝔼[βn(d)βn1(d)]\displaystyle\operatorname{\mathbb{E}{}}[\beta_{n}(d)-\beta_{n-1}(d)] =𝔼[Rn(d)Rn1(d)]𝔼rn(d)+𝔼rn1(d)\displaystyle=\operatorname{\mathbb{E}{}}[R_{n}(d)-R_{n-1}(d)]-\operatorname{\mathbb{E}{}}r_{n}(d)+\operatorname{\mathbb{E}{}}r_{n-1}(d)
=d(𝐗(n)sets a record)𝔼Rn(d1)+𝔼Rn1(d1)\displaystyle=\operatorname{\mathbb{P}{}}_{d}(\mathbf{X}^{(n)}\leavevmode\nobreak\ \text{sets a record})-\operatorname{\mathbb{E}{}}R_{n}(d-1)+\operatorname{\mathbb{E}{}}R_{n-1}(d-1)
=d(𝐗(n)sets a record)d1(𝐗(n)sets a record),\displaystyle=\operatorname{\mathbb{P}{}}_{d}(\mathbf{X}^{(n)}\leavevmode\nobreak\ \text{sets a record})-\operatorname{\mathbb{P}{}}_{d-1}(\mathbf{X}^{(n)}\leavevmode\nobreak\ \text{sets a record}),

where the second equality follows by standard consideration of concomitants: The two random variables rn(d)r_{n}(d) and Rn(d1)R_{n}(d-1) have the same distribution, for any nn and dd. Combining the last display with (3.1), one has

𝔼(Kn|Kn0)=1d1(𝐗(n)sets a record)d(𝐗(n)sets a record).\operatorname{\mathbb{E}{}}(K_{n}\,|\,K_{n}\geq 0)=1-\frac{\operatorname{\mathbb{P}{}}_{d-1}(\mathbf{X}^{(n)}\leavevmode\nobreak\ \text{sets a record})}{\operatorname{\mathbb{P}{}}_{d}(\mathbf{X}^{(n)}\leavevmode\nobreak\ \text{sets a record})}. (3.2)

Thus, by (2.1), as nn\to\infty we have

𝔼(Kn|Kn0)=1(1+o(1))d1lnn1,\operatorname{\mathbb{E}{}}(K_{n}\,|\,K_{n}\geq 0)=1-(1+o(1))\frac{d-1}{\ln n}\to 1,

as claimed. ∎

Lemma 3.3.

The sequence of probability measures μn=(KnKn0)\mu_{n}={\mathcal{L}}(K_{n}\mid K_{n}\geq 0) converges vaguely.

Proof.

It is sufficient to show that there exists (κk)k=1,2,(\kappa_{k})_{k=1,2,\dots} such that for each kk we have

(KnkKn0)κk as n.\operatorname{\mathbb{P}{}}(K_{n}\geq k\mid K_{n}\geq 0)\to\kappa_{k}\mbox{\ as $n\to\infty$}. (3.3)

This is proved by means of the next three lemmas and the arguments interspersed among them. ∎

Lemma 3.4.

Let Δ>0\Delta>0. Let Kn(Δ+)=1/2K_{n}(\Delta+)=-1/2 if 𝐗(n)\mathbf{X}^{(n)} does not set a record, and otherwise let Kn(Δ+)K_{n}(\Delta+) denote the number of remaining records killed by 𝐗(n)\mathbf{X}^{(n)} that are at 1\ell^{1}-distance >Δ>\Delta from 𝐗(n)\mathbf{X}^{(n)}. Then

lim supn(Kn(Δ+)1Kn0)\displaystyle\limsup_{n\to\infty}\operatorname{\mathbb{P}{}}(K_{n}(\Delta+)\geq 1\mid K_{n}\geq 0) limn𝔼(Kn(Δ+)Kn0)\displaystyle\leq\lim_{n\to\infty}\operatorname{\mathbb{E}{}}(K_{n}(\Delta+)\mid K_{n}\geq 0) (3.4)
=(Gamma(d)>Δ)eΔΔd1(d1)!0\displaystyle=\operatorname{\mathbb{P}{}}(\mbox{\rm Gamma$(d)$}>\Delta)\sim e^{-\Delta}\frac{\Delta^{d-1}}{(d-1)!}\to 0 (3.5)

where Gamma(d)(d) is distributed Gamma(d,1)(d,1) and the asymptotics in (3.5) are as Δ\Delta\to\infty.

Proof.

Recalling that we use \|\cdot\| to denote 1\ell^{1}-norm,

(Kn0,Kn(Δ+)1)\displaystyle\operatorname{\mathbb{P}{}}(K_{n}\geq 0,\,K_{n}(\Delta+)\geq 1)
𝔼(Kn(Δ+);Kn0)\displaystyle\leq\operatorname{\mathbb{E}{}}(K_{n}(\Delta+);K_{n}\geq 0)
=(n1)(Kn10,𝐗(n)𝐗(n1),𝐗(n)𝐗(n1)>Δ).\displaystyle=(n-1)\operatorname{\mathbb{P}{}}(K_{n-1}\geq 0,\,\mathbf{X}^{(n)}\succ\mathbf{X}^{(n-1)},\,\|\mathbf{X}^{(n)}-\mathbf{X}^{(n-1)}\|>\Delta). (3.6)

But

(Kn10,𝐗(n)𝐗(n1),𝐗(n)𝐗(n1)>Δ)\displaystyle\operatorname{\mathbb{P}{}}(K_{n-1}\geq 0,\,\mathbf{X}^{(n)}\succ\mathbf{X}^{(n-1)},\,\|\mathbf{X}^{(n)}-\mathbf{X}^{(n-1)}\|>\Delta)
=𝐳,𝐱:𝟎𝐳𝐱,𝐱𝐳>Δ(𝐗(n)d𝐱;𝐗(n1)d𝐳;\displaystyle=\int_{\mathbf{z},\mathbf{x}:{\bf 0}\prec\mathbf{z}\prec\mathbf{x},\,\|\mathbf{x}-\mathbf{z}\|>\Delta}\!\operatorname{\mathbb{P}{}}\left(\mathbf{X}^{(n)}\in\mathrm{d}\mathbf{x};\,\mathbf{X}^{(n-1)}\in\mathrm{d}\mathbf{z};\right.
𝐗(i)𝐳 for i=1,,n2)\displaystyle{}\hskip 130.08621pt\left.\mathbf{X}^{(i)}\not\succ\mathbf{z}\mbox{\ for $i=1,\dots,n-2$}\right)
=𝐳,𝐱:𝟎𝐳𝐱,𝐱𝐳>Δe𝐱e𝐳(1e𝐳)n2d𝐱d𝐳\displaystyle=\int_{\mathbf{z},\mathbf{x}:{\bf 0}\prec\mathbf{z}\prec\mathbf{x},\,\|\mathbf{x}-\mathbf{z}\|>\Delta}\!e^{-\|\mathbf{x}\|}\,e^{-\|\mathbf{z}\|}\,\left(1-e^{-\|\mathbf{z}\|}\right)^{n-2}\,\mathrm{d}\mathbf{x}\,\mathrm{d}\mathbf{z}
=𝐳𝟎𝜹𝟎:𝜹>Δe(𝐳+𝜹)e𝐳(1e𝐳)n2d𝜹d𝐳\displaystyle=\int_{\mathbf{z}\succ{\bf 0}}\int_{\bm{\delta}\succ{\bf 0}:\|\bm{\delta}\|>\Delta}\!e^{-(\|\mathbf{z}\|+\|\bm{\delta}\|)}\,e^{-\|\mathbf{z}\|}\,\left(1-e^{-\|\mathbf{z}\|}\right)^{n-2}\,\mathrm{d}\bm{\delta}\,\mathrm{d}\mathbf{z}
=[Δyd1(d1)!eydy]×𝐳𝟎e2𝐳(1e𝐳)n2d𝐳.\displaystyle=\left[\int_{\Delta}^{\infty}\!\frac{y^{d-1}}{(d-1)!}e^{-y}\,\mathrm{d}y\right]\times\int_{\mathbf{z}\succ{\bf 0}}\,e^{-2\|\mathbf{z}\|}\,\left(1-e^{-\|\mathbf{z}\|}\right)^{n-2}\,\mathrm{d}\mathbf{z}.

The first factor equals (Gamma(d)>Δ)\operatorname{\mathbb{P}{}}(\mbox{Gamma}(d)>\Delta). Utilizing all three parts of Remark 2.1, (i) the second factor, when divided by pn1p_{n-1}, equals

𝔼(e𝐗(n1)𝐗(n1) sets a record)=(n1)1𝔼eGn1;\operatorname{\mathbb{E}{}}\left(e^{-\|\mathbf{X}^{(n-1)}\|}\mid\mbox{$\mathbf{X}^{(n-1)}$ sets a record}\right)=(n-1)^{-1}\operatorname{\mathbb{E}{}}e^{-G_{n-1}};

and thus (ii) 𝔼(Kn(Δ+)Kn0)\operatorname{\mathbb{E}{}}(K_{n}(\Delta+)\mid K_{n}\geq 0) equals

(Gamma(d)>Δ)×pn1pn𝔼eGn1\displaystyle\operatorname{\mathbb{P}{}}(\mbox{Gamma}(d)>\Delta)\times\frac{p_{n-1}}{p_{n}}\operatorname{\mathbb{E}{}}e^{-G_{n-1}} (Gamma(d)>Δ)×𝔼eGn1\displaystyle\sim\operatorname{\mathbb{P}{}}(\mbox{Gamma}(d)>\Delta)\times\operatorname{\mathbb{E}{}}e^{-G_{n-1}}
(Gamma(d)>Δ),\displaystyle\to\operatorname{\mathbb{P}{}}(\mbox{Gamma}(d)>\Delta),

as nn\to\infty. The tail-asymptotics for Γ(d)\Gamma(d) appearing in (3.5) are standard. This completes the proof of the lemma. ∎

Having suitably controlled Kn(Δ+)K_{n}(\Delta+), we turn our attention to Kn(Δ)K_{n}(\Delta-), defined as follows for given Δ>0\Delta>0. Let Kn(Δ)=1/2K_{n}(\Delta-)=-1/2 if 𝐗(n)\mathbf{X}^{(n)} does not set a record, and otherwise let Kn(Δ)K_{n}(\Delta-) denote the number of remaining records killed by 𝐗(n)\mathbf{X}^{(n)} that are at 1\ell^{1}-distance Δ\leq\Delta from 𝐗(n)\mathbf{X}^{(n)}; thus

Kn=Kn(Δ)+Kn(Δ+).K_{n}=K_{n}(\Delta-)+K_{n}(\Delta+). (3.7)

It follows using Remark 2.1(b) that if we can prove, for each k{1,2,}k\in\{1,2,\ldots\}, that

g,𝐮(Gdg,𝐔du)(Kn(Δ)kKn0,Gn=g,𝐔n=𝐮)\int_{g,\mathbf{u}}\operatorname{\mathbb{P}{}}(G\in\mathrm{d}g,\,\mathbf{U}\in\mathrm{d}u)\,\operatorname{\mathbb{P}{}}(K_{n}(\Delta-)\geq k\mid K_{n}\geq 0,\,G_{n}=g,\mathbf{U}_{n}=\mathbf{u}) (3.8)

has a limit, call it κk(Δ)\kappa_{k}(\Delta-), as nn\to\infty, then, also for each kk, we have

(Kn(Δ)kKn0)κk(Δ) as n.\operatorname{\mathbb{P}{}}(K_{n}(\Delta-)\geq k\mid K_{n}\geq 0)\to\kappa_{k}(\Delta-)\mbox{\ as $n\to\infty$}. (3.9)

To prove that (3.8) has a limit, it suffices by the dominated convergence to prove the following lemma, in which case we can then take

κk(Δ):=g(Gdg)κk(Δ;,g).\kappa_{k}(\Delta-):=\int_{g}\operatorname{\mathbb{P}{}}(G\in\mathrm{d}g)\,\kappa_{k}(\Delta-;\infty,g). (3.10)
Lemma 3.5.

Let k{1,2,}k\in\{1,2,\dots\}, gg\in\mathbb{R}, and 𝐮𝒮d1\mathbf{u}\in{\mathcal{S}}_{d-1} with 𝐮𝟎\mathbf{u}\succ{\bf 0}. Then

κk(Δ;n,g,𝐮):=(Kn(Δ)kKn0,Gn=g,𝐔n=𝐮)\kappa_{k}(\Delta-;n,g,\mathbf{u}):=\operatorname{\mathbb{P}{}}(K_{n}(\Delta-)\geq k\mid K_{n}\geq 0,\,G_{n}=g,\,\mathbf{U}_{n}=\mathbf{u})

has a limit κk(Δ;,g)\kappa_{k}(\Delta-;\infty,g) as nn\to\infty, which doesn’t depend on 𝐮\mathbf{u}.

Proof.

We use the method of moments [and so, by the way, κk(Δ;,g)0\kappa_{k}(\Delta-;\infty,g)\downarrow 0 as kk\uparrow\infty; that is, the conditional distributions in question have a weak limit]. Let 𝐱=(Ln+g)𝐮\mathbf{x}=(\operatorname{L}n+g)\mathbf{u} (which implies 𝐱=Ln+g\|\mathbf{x}\|=\operatorname{L}n+g), and let RR(Δ;n1,𝐱)R\equiv R(\Delta;n-1,\mathbf{x}) denote the set of remaining record-values 𝐗(i)\mathbf{X}^{(i)} at time n1n-1 that are killed by 𝐗(n)=𝐱\mathbf{X}^{(n)}=\mathbf{x} at time nn and satisfy 𝐱𝐗(i)Δ\|\mathbf{x}-\mathbf{X}^{(i)}\|\leq\Delta. By writing Kn(Δ)K_{n}(\Delta-) as a sum of indicators, for integer r1r\geq 1 we calculate

𝔼[(Kn(Δ))rKn0,Gn=g,𝐔n=𝐮]\displaystyle\operatorname{\mathbb{E}{}}[(K_{n}(\Delta-))^{r}\mid K_{n}\geq 0,\,G_{n}=g,\,\mathbf{U}_{n}=\mathbf{u}] (3.11)
=m=1r(rj1,,jm)({𝐗(i1),,𝐗(im)}R𝐗(n)=𝐱 sets a record)\displaystyle=\sum_{m=1}^{r}\sum\sum{r\choose{j_{1},\dots,j_{m}}}\operatorname{\mathbb{P}{}}(\{\mathbf{X}^{(i_{1})},\ldots,\mathbf{X}^{(i_{m})}\}\subseteq R\mid\mbox{$\mathbf{X}^{(n)}=\mathbf{x}$ sets a record}) (3.12)
=m=1r(n1m)m!{rm}({𝐗(1),,𝐗(m)}R𝐗(n)=𝐱 sets a record),\displaystyle=\sum_{m=1}^{r}{{n-1}\choose m}m!{r\brace m}\operatorname{\mathbb{P}{}}(\{\mathbf{X}^{(1)},\ldots,\mathbf{X}^{(m)}\}\subseteq R\mid\mbox{$\mathbf{X}^{(n)}=\mathbf{x}$ sets a record}), (3.13)

where, in (3.12), the second of the three sums is over i1,,imi_{1},\dots,i_{m} satisfying 1i1<<imn11\leq i_{1}<\cdots<i_{m}\leq n-1 and the third sum is over j1,,jm1j_{1},\dots,j_{m}\geq 1 satisfying j+=rj_{+}=r; and, in (3.13), {rm}{r\brace m} is a Stirling number of the second kind and is the number of ways to partition an rr-set into mm nonempty sets.

Now, writing 𝐗=𝐗(1)\mathbf{X}=\mathbf{X}^{(1)}, the conditional probability in (3.13) equals

1[(𝐗𝐱)]n1[i=1m(e𝐱(i)d𝐱(i))][(i=1m{𝐗𝐱(i)})]nm,\frac{1}{[\operatorname{\mathbb{P}{}}(\mathbf{X}\not\succ\mathbf{x})]^{n-1}}\int\!\left[\prod_{i=1}^{m}\left(e^{-\|\mathbf{x}^{(i)}\|}\,\mathrm{d}\mathbf{x}^{(i)}\right)\right]\left[\operatorname{\mathbb{P}{}}\left(\bigcap_{i=1}^{m}\left\{\mathbf{X}\not\succ\mathbf{x}^{(i)}\right\}\right)\right]^{n-m},

where the integral is over incomparable vectors 𝐱(1),,𝐱(m)\mathbf{x}^{(1)},\dots,\mathbf{x}^{(m)} such that, for i=1,,mi=1,\dots,m, we have 𝟎𝐱(i)𝐱{\bf 0}\prec\mathbf{x}^{(i)}\prec\mathbf{x} and 𝐱𝐱(i)Δ\|\mathbf{x}-\mathbf{x}^{(i)}\|\leq\Delta. For the denominator here we calculate

[(𝐗𝐱)]n1=(1e𝐱)n1=(1n1eg)n1exp(eg).[\operatorname{\mathbb{P}{}}(\mathbf{X}\not\succ\mathbf{x})]^{n-1}=\left(1-e^{-\|\mathbf{x}\|}\right)^{n-1}=\left(1-n^{-1}e^{-g}\right)^{n-1}\to\exp\left(-e^{-g}\right).

Changing variables from 𝐱(i)\mathbf{x}^{(i)} to 𝜹(i)=𝐱𝐱(i)\bm{\delta}^{(i)}=\mathbf{x}-\mathbf{x}^{(i)}, the numerator (i.e., integral) can be written as

em𝐱[i=1m(e𝜹(i)d𝜹(i))][(i=1m{𝐗𝐱𝜹(i)})]nm\displaystyle\int\!e^{-m\|\mathbf{x}\|}\left[\prod_{i=1}^{m}\left(e^{\|\bm{\delta}^{(i)}\|}\,\mathrm{d}\bm{\delta}^{(i)}\right)\right]\left[\operatorname{\mathbb{P}{}}\left(\bigcap_{i=1}^{m}\left\{\mathbf{X}\not\succ\mathbf{x}-\bm{\delta}^{(i)}\right\}\right)\right]^{n-m} (3.14)
=nmemgA[i=1m(e𝜹(i)d𝜹(i))][(i=1m{𝐗𝐱𝜹(i)})]nm.\displaystyle=n^{-m}e^{-mg}\int\!A\left[\prod_{i=1}^{m}\left(e^{\|\bm{\delta}^{(i)}\|}\,\mathrm{d}\bm{\delta}^{(i)}\right)\right]\left[\operatorname{\mathbb{P}{}}\left(\bigcap_{i=1}^{m}\left\{\mathbf{X}\not\succ\mathbf{x}-\bm{\delta}^{(i)}\right\}\right)\right]^{n-m}. (3.15)

Both integrals are over vectors 𝜹(1),,𝜹(m)\bm{\delta}^{(1)},\dots,\bm{\delta}^{(m)} satisfying the restrictions that

𝜹(1),,𝜹(m)\bm{\delta}^{(1)},\dots,\bm{\delta}^{(m)} are incomparable, and, (3.16)
for i=1,,m, we have 𝟎𝜹(i) and 𝜹(i)Δ;\displaystyle{}\qquad\quad\mbox{for $i=1,\dots,m$, we have ${\bf 0}\prec\bm{\delta}^{(i)}$ and $\|\bm{\delta}^{(i)}\|\leq\Delta$};

for the integral in (3.14) we have the additional restrictions that 𝜹(i)𝐱\bm{\delta}^{(i)}\prec\mathbf{x}, and, correspondingly, in the second integral we use the shorthand

A=𝟏(𝜹(i)𝐱 for i=1,,m).A={\bf 1}(\bm{\delta}^{(i)}\prec\mathbf{x}\mbox{\ for $i=1,\dots,m$}).

Observe that the integrands in (3.15) can be dominated by

i=1me𝜹(i),\prod_{i=1}^{m}e^{\|\bm{\delta}^{(i)}\|}, (3.17)

which is integrable [over the specified range (3.16) of 𝜹(1),,𝜹(m)\bm{\delta}^{(1)},\dots,\bm{\delta}^{(m)}] because, dropping the restriction of incomparability in the next integral to appear, the integral of (3.17) can be bounded by

i=1m(e𝜹(i)d𝜹(i))\displaystyle\int\!\prod_{i=1}^{m}\left(e^{\|\bm{\delta}^{(i)}\|}\,\mathrm{d}\bm{\delta}^{(i)}\right) =[𝜹𝟎:𝜹Δe𝜹d𝜹]m\displaystyle=\left[\int_{\bm{\delta}\succ{\bf 0}:\|\bm{\delta}\|\leq\Delta}\!e^{\|\bm{\delta}\|}\,\mathrm{d}\bm{\delta}\right]^{m}
=[0Δyd1(d1)!eydy]m\displaystyle=\left[\int_{0}^{\Delta}\!\frac{y^{d-1}}{(d-1)!}\,e^{y}\,\mathrm{d}y\right]^{m}
[0Δe2ydy]m\displaystyle\leq\left[\int_{0}^{\Delta}\!e^{2y}\,\mathrm{d}y\right]^{m}
(12e2Δ)m=2me2Δm<.\displaystyle\leq\left(\frac{1}{2}e^{2\Delta}\right)^{m}=2^{-m}e^{2\Delta m}<\infty. (3.18)

So we may apply the dominated convergence theorem to the integral appearing in (3.15). The assumption 𝐮𝟎\mathbf{u}\succ{\bf 0} of strict positivity is crucial, since it implies that A1A\to 1 as nn\to\infty. If we can show, for fixed mm and gg and 𝐮\mathbf{u} and 𝜹(1),𝜹(m)\bm{\delta}^{(1)},\ldots\bm{\delta}^{(m)} that

qn\displaystyle q_{n} qn(m,g,𝐮,𝜹(1),,𝜹(m)):=[(i=1m{𝐗𝐱𝜹(i)})]nm\displaystyle\equiv q_{n}(m,g,\mathbf{u},\bm{\delta}^{(1)},\dots,\bm{\delta}^{(m)}):=\left[\operatorname{\mathbb{P}{}}\left(\bigcap_{i=1}^{m}\left\{\mathbf{X}\not\succ\mathbf{x}-\bm{\delta}^{(i)}\right\}\right)\right]^{n-m}
has a limit q(m,g,𝐮,𝜹(1),,𝜹(m)) as n,\displaystyle{}\quad\mbox{has a limit $q_{\infty}(m,g,\mathbf{u},\bm{\delta}^{(1)},\dots,\bm{\delta}^{(m)})$ as $n\to\infty$}, (3.19)

then (3.11) has the following limit as nn\to\infty:

exp(eg)m=1r{rm}emg\displaystyle\exp\!\left(e^{-g}\right)\sum_{m=1}^{r}{r\brace m}e^{-mg}
×[i=1m(e𝜹(i)d𝜹(i))]q(m,g,𝐮,𝜹(1),,𝜹(m)),\displaystyle\times\int\!\left[\prod_{i=1}^{m}\left(e^{\|\bm{\delta}^{(i)}\|}\,\mathrm{d}\bm{\delta}^{(i)}\right)\right]q_{\infty}(m,g,\mathbf{u},\bm{\delta}^{(1)},\dots,\bm{\delta}^{(m)}), (3.20)

where the integral is over 𝜹(1),𝜹(m)\bm{\delta}^{(1)},\ldots\bm{\delta}^{(m)} satisfying (3.16). Further, using (3.1) this limit is bounded by

exp(eg)m=1r{rm}emg2me2Δm\displaystyle\exp\!\left(e^{-g}\right)\sum_{m=1}^{r}{r\brace m}e^{-mg}2^{-m}e^{2\Delta m} exp(eg)m=1rmrm!emg2me2Δm\displaystyle\leq\exp\!\left(e^{-g}\right)\sum_{m=1}^{r}\frac{m^{r}}{m!}e^{-mg}2^{-m}e^{2\Delta m}
12exp(eg)rr+1er|g|e2Δr.\displaystyle\leq\frac{1}{2}\exp\!\left(e^{-g}\right)r^{r+1}e^{r|g|}e^{2\Delta r}.

Given the rate of growth of this bound as a function of rr, the method of moments applies: The conditional distribution of Kn(Δ)K_{n}(\Delta-) given Kn0K_{n}\geq 0, Gn=gG_{n}=g, and 𝐔n=𝐮\mathbf{U}_{n}=\mathbf{u} converges weakly to the unique probability measure on the nonnegative integers whose rrth moment is given by (3.20) for r=1,2,r=1,2,\dots.

It remains to prove (3.19). Indeed, defining 𝜹(i1,,i)\bm{\delta}^{(i_{1},\ldots,i_{\ell})} to be the coordinate-wise minimum h=1𝜹(ih)\wedge_{h=1}^{\ell}\bm{\delta}^{(i_{h})} and applying inclusion–exclusion at the second equality,

qn\displaystyle q_{n} =[1(i=1m{𝐗𝐱𝜹(i)})]nm\displaystyle=\left[1-\operatorname{\mathbb{P}{}}\left(\bigcup_{i=1}^{m}\left\{\mathbf{X}\succ\mathbf{x}-\bm{\delta}^{(i)}\right\}\right)\right]^{n-m}
=[1=1m(1)11i1<i2<<im(h=1{𝐗𝐱𝜹(ih)})]nm\displaystyle=\left[1-\sum_{\ell=1}^{m}(-1)^{\ell-1}\sum_{1\leq i_{1}<i_{2}<\cdots<i_{\ell}\leq m}\operatorname{\mathbb{P}{}}\left(\bigcap_{h=1}^{\ell}\left\{\mathbf{X}\succ\mathbf{x}-\bm{\delta}^{(i_{h})}\right\}\right)\right]^{n-m}
=[1=1m(1)11i1<i2<<im(𝐗𝐱𝜹(i1,,i)))]nm\displaystyle=\left[1-\sum_{\ell=1}^{m}(-1)^{\ell-1}\sum_{1\leq i_{1}<i_{2}<\cdots<i_{\ell}\leq m}\operatorname{\mathbb{P}{}}\left(\mathbf{X}\succ\mathbf{x}-\bm{\delta}^{(i_{1},\ldots,i_{\ell}))}\right)\right]^{n-m}
=[1e𝐱=1m(1)11i1<i2<<ime𝜹(i1,,i))]nm\displaystyle=\left[1-e^{-\|\mathbf{x}\|}\sum_{\ell=1}^{m}(-1)^{\ell-1}\sum_{1\leq i_{1}<i_{2}<\cdots<i_{\ell}\leq m}e^{\|\bm{\delta}^{(i_{1},\ldots,i_{\ell}))}\|}\right]^{n-m}
=[1n1eg=1m(1)11i1<i2<<ime𝜹(i1,,i))]nm\displaystyle=\left[1-n^{-1}e^{-g}\sum_{\ell=1}^{m}(-1)^{\ell-1}\sum_{1\leq i_{1}<i_{2}<\cdots<i_{\ell}\leq m}e^{\|\bm{\delta}^{(i_{1},\ldots,i_{\ell}))}\|}\right]^{n-m}
exp[eg=1m(1)11i1<i2<<ime𝜹(i1,,i))] as n\displaystyle\longrightarrow\exp\left[-e^{-g}\sum_{\ell=1}^{m}(-1)^{\ell-1}\sum_{1\leq i_{1}<i_{2}<\cdots<i_{\ell}\leq m}e^{\|\bm{\delta}^{(i_{1},\ldots,i_{\ell}))}\|}\right]\mbox{\ as $n\to\infty$}
=:q(m,g,𝜹(1),,𝜹(m)),\displaystyle=:q_{\infty}(m,g,\bm{\delta}^{(1)},\dots,\bm{\delta}^{(m)}), (3.21)

as desired. ∎

Since the expression κk(Δ;n,g,𝐮)\kappa_{k}(\Delta;n,g,\mathbf{u}) studied in Lemma 3.5 is clearly nondecreasing in Δ\Delta, the same is true for its limit κk(Δ;,g)\kappa_{k}(\Delta;\infty,g)—and therefore, by (3.10), for the limit κk(Δ)\kappa_{k}(\Delta-) appearing in (3.9).

Lemma 3.6.

The convergence

(KnkKn0)κk as n\operatorname{\mathbb{P}{}}(K_{n}\geq k\mid K_{n}\geq 0)\to\kappa_{k}\mbox{\ as $n\to\infty$}

claimed at (3.3) holds with

κk:=limΔκk(Δ),\kappa_{k}:=\lim_{\Delta\to\infty}\kappa_{k}(\Delta-), (3.22)

where κk(Δ)\kappa_{k}(\Delta-) is defined at (3.10) using Lemma 3.5.

Proof.

This follows simply from (3.7), Lemma 3.4, and (3.9). Indeed, the conditional probability in question is, for each Δ\Delta, at least as large as the one in (3.9) and so, by (3.9), has a lim inf\liminf at least κk(Δ)\kappa_{k}(\Delta-). Letting Δ\Delta\to\infty, we find

lim infn(KnkKn0)κk.\liminf_{n\to\infty}\operatorname{\mathbb{P}{}}(K_{n}\geq k\mid K_{n}\geq 0)\geq\kappa_{k}.

Conversely, the conditional probability in question is, by finite subadditivity, at most the sum of the one in (3.9) and the one treated in Lemma 3.4. Thus, for each Δ\Delta, we have

lim supn(KnkKn0)κk(Δ)+(Gamma(d)>Δ)\limsup_{n\to\infty}\operatorname{\mathbb{P}{}}(K_{n}\geq k\mid K_{n}\geq 0)\leq\kappa_{k}(\Delta-)+\operatorname{\mathbb{P}{}}(\mbox{Gamma}(d)>\Delta)

by (3.9) and Lemma 3.4. Letting Δ\Delta\to\infty, we find from (3.22) and (3.5) that

lim supn(KnkKn0)κk.\limsup_{n\to\infty}\operatorname{\mathbb{P}{}}(K_{n}\geq k\mid K_{n}\geq 0)\leq\kappa_{k}.

This completes the proof. ∎

3.2. Identification of the weak limit μ\mu as (𝒦){\mathcal{L}}(\mathcal{K})

In this subsection we prove the following proposition. Recall that the existence of a weak limit μ\mu for the probability measures μn=(KnKn0)\mu_{n}={\mathcal{L}}(K_{n}\mid K_{n}\geq 0) is established in Proposition 3.1 and identified to some extent in Lemma 3.6.

Proposition 3.7.

The weak limit μ\mu of the measures μn=(KnKn0)\mu_{n}={\mathcal{L}}(K_{n}\mid K_{n}\geq 0) is the distribution (𝒦){\mathcal{L}}(\mathcal{K}) described in Theorem 1.3.

Our approach to proving this proposition is to define 𝒦(Δ)\mathcal{K}(\Delta-) and 𝒦(Δ+)\mathcal{K}(\Delta+) in relation to 𝒦\mathcal{K} in the same way that we defined Kn(Δ)K_{n}(\Delta-) and Kn(Δ+)K_{n}(\Delta+) in relation to KnK_{n}, to prove an analogue (namely, Lemma 3.8) of Lemma 3.4, and to use again the method of moments to establish (see Lemma 3.9) that the limit κk(Δ;,g)\kappa_{k}(\Delta-;\infty,g) in Lemma 3.5 equals (𝒦g(Δ)k)\operatorname{\mathbb{P}{}}(\mathcal{K}_{g}(\Delta-)\geq k) (in notation that should be obvious and which we shall at any rate make explicit in the statement of Lemma 3.9).

After stating and proving Lemmas 3.83.9, we will give the (then) easy proof of Proposition 3.7.

Lemma 3.8.

Let Δ>0\Delta>0. For gg\in\mathbb{R}, let

𝒦g(Δ+)\displaystyle\mathcal{K}_{g}(\Delta+) =number of maxima dominated by 𝐱𝐱(g):=(g/d,,g/d)\displaystyle=\mbox{\rm number of maxima dominated by $\mathbf{x}\equiv\mathbf{x}(g):=(g/d,\ldots,g/d)$}
and at 1\ell^{1}-distance >Δ>\Delta from 𝐱\mathbf{x}

in the nonhomogeneous Poisson point process described in Theorem 1.3, with intensity function

𝟏(𝐳𝐱)ez+,𝐳d.{\bf 1}(\mathbf{z}\not\succ\mathbf{x})e^{-z_{+}},\quad\mathbf{z}\in\mathbb{R}^{d}.

Define (𝒦(Δ+)){\mathcal{L}}(\mathcal{K}(\Delta+)) to be the mixture 𝔼(𝒦G(Δ+))\operatorname{\mathbb{E}{}}{\mathcal{L}}(\mathcal{K}_{G}(\Delta+)), where GG is distributed standard Gumbel. Then

(𝒦(Δ+)1)𝔼𝒦(Δ+)=(Gamma(d)>Δ)eΔΔd1(d1)!0\operatorname{\mathbb{P}{}}(\mathcal{K}(\Delta+)\geq 1)\leq\operatorname{\mathbb{E}{}}\mathcal{K}(\Delta+)=\operatorname{\mathbb{P}{}}(\mbox{\rm Gamma}(d)>\Delta)\sim e^{-\Delta}\frac{\Delta^{d-1}}{(d-1)!}\to 0

where Gamma(d)\mbox{\rm Gamma}(d) is distributed Gamma(d,1)(d,1) and the asymptotics are as Δ\Delta\to\infty.

Proof.

Let

𝒫:={𝐲d:𝐲𝐱,𝐱𝐲>Δ}.\mathcal{P}:=\{\mathbf{y}\in\mathbb{R}^{d}:\mathbf{y}\prec\mathbf{x},\,\|\mathbf{x}-\mathbf{y}\|>\Delta\}.

Denote the Poisson process by NgN_{g}. Apply the so-called Mecke equation by setting

f(𝐳,η)=𝟏(𝐳 is a maximum of η dominated by 𝐱 and at 1-distance from 𝐱)f(\mathbf{z},\eta)={\bf 1}(\mbox{$\mathbf{z}$ is a maximum of\leavevmode\nobreak\ $\eta$ dominated by\leavevmode\nobreak\ $\mathbf{x}$ and at $\ell^{1}$-distance from\leavevmode\nobreak\ $\mathbf{x}$})

and taking η=Ng\eta=N_{g} in [6, Theorem 4.1] to find

𝔼𝒦g(Δ+)\displaystyle\operatorname{\mathbb{E}{}}\mathcal{K}_{g}(\Delta+)
=𝐲𝒫ey+exp(𝐳:𝐳𝐲,𝐳𝐱ez+d𝐳)d𝐲\displaystyle=\int_{\mathbf{y}\in\mathcal{P}}e^{-y_{+}}\,\exp\!\left(-\int_{\mathbf{z}:\mathbf{z}\succ\mathbf{y},\,\mathbf{z}\not\succ\mathbf{x}}\!e^{-z_{+}}\,\mathrm{d}\mathbf{z}\right)\,\mathrm{d}\mathbf{y}
=𝐲𝒫ey+exp[(𝐳:𝐳𝐲ez+d𝐳𝐳:𝐳𝐱ez+d𝐳)]d𝐲\displaystyle=\int_{\mathbf{y}\in\mathcal{P}}e^{-y_{+}}\,\exp\!\left[-\left(\int_{\mathbf{z}:\mathbf{z}\succ\mathbf{y}}\!e^{-z_{+}}\,\mathrm{d}\mathbf{z}-\int_{\mathbf{z}:\mathbf{z}\succ\mathbf{x}}\!e^{-z_{+}}\,\mathrm{d}\mathbf{z}\right)\right]\,\mathrm{d}\mathbf{y}
=𝐲𝒫ey+exp[(ey+ex+)]d𝐲\displaystyle=\int_{\mathbf{y}\in\mathcal{P}}e^{-y_{+}}\,\exp\!\left[-\left(e^{-y_{+}}-e^{-x_{+}}\right)\right]\,\mathrm{d}\mathbf{y}
=exp(eg)𝐲𝒫ey+exp(ey+)d𝐲\displaystyle=\exp\!\left(e^{-g}\right)\int_{\mathbf{y}\in\mathcal{P}}\!e^{-y_{+}}\,\exp\!\left(-e^{-y_{+}}\right)\,\mathrm{d}\mathbf{y}
=exp(eg)𝜹𝟎:𝜹>Δe(x+𝜹)exp[e(x+𝜹)]d𝜹\displaystyle=\exp\!\left(e^{-g}\right)\int_{\bm{\delta}\succ{\bf 0}:\|\bm{\delta}\|>\Delta}\!e^{-\left(x_{+}-\|\bm{\delta}\|\right)}\,\exp\!\left[-e^{-\left(x_{+}-\|\bm{\delta}\|\right)}\right]\,\mathrm{d}\bm{\delta}
=exp(eg)𝜹𝟎:𝜹>Δe(g𝜹)exp[e(g𝜹)]d𝜹\displaystyle=\exp\!\left(e^{-g}\right)\int_{\bm{\delta}\succ{\bf 0}:\|\bm{\delta}\|>\Delta}\!e^{-\left(g-\|\bm{\delta}\|\right)}\,\exp\!\left[-e^{-\left(g-\|\bm{\delta}\|\right)}\right]\,\mathrm{d}\bm{\delta}
=exp(eg)Δηd1(d1)!e(gη)exp[e(gη)]dη.\displaystyle=\exp\!\left(e^{-g}\right)\int_{\Delta}^{\infty}\!\frac{\eta^{d-1}}{(d-1)!}\,e^{-(g-\eta)}\,\exp\!\left[-e^{-(g-\eta)}\right]\,\mathrm{d}\eta.

Multiply by (Gdg)=exp(eg)egdg\operatorname{\mathbb{P}{}}(G\in\mathrm{d}g)=\exp\!\left(-e^{-g}\right)e^{-g}\,\mathrm{d}g and integrate over gg\in\mathbb{R} to get

(𝒦(Δ+)1)\displaystyle\operatorname{\mathbb{P}{}}(\mathcal{K}(\Delta+)\geq 1) 𝔼𝒦(Δ+)\displaystyle\leq\operatorname{\mathbb{E}{}}\mathcal{K}(\Delta+)
=Δηd1(d1)!eηe2gexp(eηeg)dgdη\displaystyle=\int_{\Delta}^{\infty}\!\frac{\eta^{d-1}}{(d-1)!}\,e^{\eta}\int_{-\infty}^{\infty}\!e^{-2g}\,\exp\!\left(-e^{\eta}e^{-g}\right)\,\mathrm{d}g\,\mathrm{d}\eta
=Δηd1(d1)!eηdη=(Gamma(d)>Δ),\displaystyle=\int_{\Delta}^{\infty}\!\frac{\eta^{d-1}}{(d-1)!}\,e^{-\eta}\,\mathrm{d}\eta=\operatorname{\mathbb{P}{}}(\mbox{Gamma}(d)>\Delta),

as claimed. ∎

Lemma 3.9.

Let Δ>0\Delta>0. For gg\in\mathbb{R}, let

𝒦g(Δ)\displaystyle\mathcal{K}_{g}(\Delta-) =number of maxima dominated by 𝐱𝐱(g):=(g/d,,g/d)\displaystyle=\mbox{\rm number of maxima dominated by $\mathbf{x}\equiv\mathbf{x}(g):=(g/d,\ldots,g/d)$}
and at 1\ell^{1}-distance Δ\leq\Delta from 𝐱\mathbf{x}

in the nonhomogeneous Poisson point process described in Theorem 1.3, with intensity function

𝟏(𝐳𝐱)ez+,𝐳d.{\bf 1}(\mathbf{z}\not\succ\mathbf{x})e^{-z_{+}},\quad\mathbf{z}\in\mathbb{R}^{d}.

Then for every k{1,2,}k\in\{1,2,\dots\} we have

(𝒦g(Δ)k)=κk(Δ;,g)\operatorname{\mathbb{P}{}}(\mathcal{K}_{g}(\Delta-)\geq k)=\kappa_{k}(\Delta-;\infty,g)

with κk(Δ;,g)\kappa_{k}(\Delta-;\infty,g) as in Lemma 3.5.

Proof.

We need only show that, for each r=1,2,r=1,2,\dots, the random variable 𝒦g(Δ)\mathcal{K}_{g}(\Delta-) has rthr^{\rm\scriptsize th} moment equal to (3.20), where q(m,g,𝜹(1),,𝜹(m))q_{\infty}(m,g,\bm{\delta}^{(1)},\dots,\bm{\delta}^{(m)}) is defined at (3.21).

For this, we proceed in much the same way as for the proof of Lemma 3.8. In this case, let

𝒫:={𝐲d:𝐲𝐱,𝐱𝐲Δ}.\mathcal{P}:=\{\mathbf{y}\in\mathbb{R}^{d}:\mathbf{y}\prec\mathbf{x},\,\|\mathbf{x}-\mathbf{y}\|\leq\Delta\}.

Applying the multivariate Mecke equation [6, Theorem 4.4], we find

𝔼[(𝒦g(Δ))r]\displaystyle\operatorname{\mathbb{E}{}}\left[\left(\mathcal{K}_{g}(\Delta-)\right)^{r}\right]
=m=1r{rm}[h=1m(ex+(h)d𝐱(h))]exp(𝐳:𝐳𝐱(h) for some h,𝐳𝐱ez+d𝐳)\displaystyle=\sum_{m=1}^{r}{r\brace m}\int\!\left[\prod_{h=1}^{m}\left(e^{-x^{(h)}_{+}}\,\mathrm{d}\mathbf{x}^{(h)}\right)\right]\exp\!\left(-\int_{\begin{subarray}{c}\mathbf{z}:\mathbf{z}\succ\mathbf{x}^{(h)}\mbox{\scriptsize\ for some\leavevmode\nobreak\ $h$},\\ \mathbf{z}\not\succ\mathbf{x}\end{subarray}}e^{-z_{+}}\,\mathrm{d}\mathbf{z}\right)
=exp(eg)m=1r{rm}[h=1m(ex+(h)d𝐱(h))]exp(𝐳:𝐳𝐱(h) for some hez+d𝐳),\displaystyle=\exp\!\left(e^{-g}\right)\sum_{m=1}^{r}{r\brace m}\int\!\left[\prod_{h=1}^{m}\left(e^{-x^{(h)}_{+}}\,\mathrm{d}\mathbf{x}^{(h)}\right)\right]\exp\!\left(-\int_{\mathbf{z}:\mathbf{z}\succ\mathbf{x}^{(h)}\mbox{\scriptsize\ for some\leavevmode\nobreak\ $h$}}e^{-z_{+}}\,\mathrm{d}\mathbf{z}\right),

where the unlabeled integrals are over incomparable vectors 𝐱(1),,𝐱(m)\mathbf{x}^{(1)},\dots,\mathbf{x}^{(m)} belonging to 𝒫\mathcal{P}.

Changing variables from 𝐱(h)\mathbf{x}^{(h)} to 𝜹(h)=𝐱𝐱(h)\bm{\delta}^{(h)}=\mathbf{x}-\mathbf{x}^{(h)}, we find

𝔼[(𝒦g(Δ))r]\displaystyle\operatorname{\mathbb{E}{}}\left[\left(\mathcal{K}_{g}(\Delta-)\right)^{r}\right]
=exp(eg)m=1r{rm}emg\displaystyle=\exp\!\left(e^{-g}\right)\sum_{m=1}^{r}{r\brace m}e^{-mg}
×[i=1m(e𝜹(i)d𝜹(i))]exp(𝐳:𝐳𝐱𝜹(i) for some iez+d𝐳),\displaystyle{}\qquad\times\int\!\left[\prod_{i=1}^{m}\left(e^{\|\bm{\delta}^{(i)}\|}\,\mathrm{d}\bm{\delta}^{(i)}\right)\right]\exp\!\left(-\int_{\mathbf{z}:\mathbf{z}\succ\mathbf{x}-\bm{\delta}^{(i)}\mbox{\scriptsize\ for some\leavevmode\nobreak\ $i$}}e^{-z_{+}}\,\mathrm{d}\mathbf{z}\right), (3.23)

where now the unlabeled integral is over vectors 𝜹(1),,𝜹(m)\bm{\delta}^{(1)},\dots,\bm{\delta}^{(m)} satisfying the restrictions (3.16). By using inclusion–exclusion, the 𝐳\mathbf{z}-integral equals [writing 𝜹(i1,,i)\bm{\delta}^{(i_{1},\dots,i_{\ell})} for the coordinate-wise minimum h=1𝜹(ih)\wedge_{h=1}^{\ell}\bm{\delta}^{(i_{h})} as in the proof of Lemma 3.5]

=1m(1)11i1<i2<<im𝐳:𝐳𝐱𝜹(ij) for all 1jez+d𝐳\displaystyle\sum_{\ell=1}^{m}(-1)^{\ell-1}\sum_{1\leq i_{1}<i_{2}<\dots<i_{\ell}\leq m}\int_{\mathbf{z}:\mathbf{z}\succ\mathbf{x}-\bm{\delta}^{(i_{j})}\mbox{\scriptsize\ for all $1\leq j\leq\ell$}}e^{-z_{+}}\,\mathrm{d}\mathbf{z}
==1m(1)11i1<i2<<im𝐳:𝐳𝐱𝜹(i1,,i)ez+d𝐳\displaystyle=\sum_{\ell=1}^{m}(-1)^{\ell-1}\sum_{1\leq i_{1}<i_{2}<\dots<i_{\ell}\leq m}\int_{\mathbf{z}:\mathbf{z}\succ\mathbf{x}-\bm{\delta}^{(i_{1},\dots,i_{\ell})}}e^{-z_{+}}\,\mathrm{d}\mathbf{z}
=eg=1m(1)11i1<i2<<ime𝜹(i1,,il).\displaystyle=e^{-g}\sum_{\ell=1}^{m}(-1)^{\ell-1}\sum_{1\leq i_{1}<i_{2}<\dots<i_{\ell}\leq m}e^{\|\bm{\delta}^{(i_{1},\dots,i_{l})}\|}.

Thus [confer (3.21)]

𝔼[(𝒦g(Δ))r]\displaystyle\operatorname{\mathbb{E}{}}\left[\left(\mathcal{K}_{g}(\Delta-)\right)^{r}\right]
=exp(eg)m=1r{rm}emg\displaystyle=\exp\!\left(e^{-g}\right)\sum_{m=1}^{r}{r\brace m}e^{-mg} (3.24)
×[i=1m(e𝜹(i)d𝜹(i))]exp(eg=1m(1)11i1<i2<<ime𝜹(i1,,il))\displaystyle{}\times\int\!\left[\prod_{i=1}^{m}\left(e^{\|\bm{\delta}^{(i)}\|}\,\mathrm{d}\bm{\delta}^{(i)}\right)\right]\exp\!\left(-e^{-g}\sum_{\ell=1}^{m}(-1)^{\ell-1}\sum_{1\leq i_{1}<i_{2}<\dots<i_{\ell}\leq m}e^{\|\bm{\delta}^{(i_{1},\dots,i_{l})}\|}\right)
=exp(eg)m=1r{rm}emg[i=1m(e𝜹(i)d𝜹(i))]q(m,g,𝜹(1),,𝜹(m))\displaystyle=\exp\!\left(e^{-g}\right)\sum_{m=1}^{r}{r\brace m}e^{-mg}\int\!\left[\prod_{i=1}^{m}\left(e^{\|\bm{\delta}^{(i)}\|}\,\mathrm{d}\bm{\delta}^{(i)}\right)\right]q_{\infty}(m,g,\bm{\delta}^{(1)},\dots,\bm{\delta}^{(m)})
=(3.20),\displaystyle=\eqref{lim moment r},

as claimed. ∎

Proof of Proposition 3.7.

Clearly,

(𝒦(Δ)k)(𝒦k)(𝒦(Δ)k)+(𝒦(Δ+)1).\operatorname{\mathbb{P}{}}(\mathcal{K}(\Delta-)\geq k)\leq\operatorname{\mathbb{P}{}}(\mathcal{K}\geq k)\leq\operatorname{\mathbb{P}{}}(\mathcal{K}(\Delta-)\geq k)+\operatorname{\mathbb{P}{}}(\mathcal{K}(\Delta+)\geq 1). (3.25)

But

(𝒦(Δ)k)\displaystyle\operatorname{\mathbb{P}{}}(\mathcal{K}(\Delta-)\geq k) =(Gdg)(𝒦g(Δ)k)\displaystyle=\int_{-\infty}^{\infty}\operatorname{\mathbb{P}{}}(G\in\,\mathrm{d}g)\operatorname{\mathbb{P}{}}(\mathcal{K}_{g}(\Delta-)\geq k)
=(Gdg)κk(Δ;,g)\displaystyle=\int_{-\infty}^{\infty}\operatorname{\mathbb{P}{}}(G\in\,\mathrm{d}g)\,\kappa_{k}(\Delta-;\infty,g) (3.26)
=κk(Δ),\displaystyle=\kappa_{k}(\Delta-), (3.27)

using Lemma 3.9 at (3.26) and (3.10) at (3.27).

Therefore, passing to the limit in (3.25) using (3.22) and recalling Lemma 3.6, we find

(𝒦k)=κk=limn(KnkKn0)=μ({k,k+1,}),\operatorname{\mathbb{P}{}}(\mathcal{K}\geq k)=\kappa_{k}=\lim_{n\to\infty}\operatorname{\mathbb{P}{}}(K_{n}\geq k\mid K_{n}\geq 0)=\mu(\{k,k+1,\dots\}),

as claimed. ∎

4. Tail probabilities for 𝒦\mathcal{K}

In Subsection 4.1 we show that all the moments of (𝒦){\mathcal{L}}(\mathcal{K}) are finite and give a simple upper bound on each, and in Subsection 4.2 we study right-tail (logarithmic) asymptotics for (𝒦){\mathcal{L}}(\mathcal{K}).

4.1. Bounds on the moments of (𝒦){\mathcal{L}}(\mathcal{K})

Here is the main result of this subsection.

Proposition 4.1.

Let 𝒦\mathcal{K} be as described in Theorem 1.3. Then for r=1,2,r=1,2,\dots we have

𝔼𝒦radrr(d+21d1)r<,\operatorname{\mathbb{E}{}}\mathcal{K}^{r}\leq a_{d}^{r}r^{(d+2-\frac{1}{d-1})r}<\infty,

with ad:=21/dd(d+1)/(d1)a_{d}:=2^{1/d}d^{(d+1)/(d-1)}.

Proof.

Recall from (3.23) that

𝔼[(𝒦g(Δ))r]\displaystyle\operatorname{\mathbb{E}{}}\left[\left(\mathcal{K}_{g}(\Delta-)\right)^{r}\right]
=exp(eg)m=1r{rm}emg\displaystyle=\exp\!\left(e^{-g}\right)\sum_{m=1}^{r}{r\brace m}e^{-mg}
×[i=1m(e𝜹(i)d𝜹(i))]exp(𝐳:𝐳𝐱𝜹(i) for some iez+d𝐳)\displaystyle{}\qquad\times\int\!\left[\prod_{i=1}^{m}\left(e^{\|\bm{\delta}^{(i)}\|}\,\mathrm{d}\bm{\delta}^{(i)}\right)\right]\exp\!\left(-\int_{\mathbf{z}:\mathbf{z}\succ\mathbf{x}-\bm{\delta}^{(i)}\mbox{\scriptsize\ for some\leavevmode\nobreak\ $i$}}e^{-z_{+}}\,\mathrm{d}\mathbf{z}\right)
=exp(eg)m=1r{rm}emg\displaystyle=\exp\!\left(e^{-g}\right)\sum_{m=1}^{r}{r\brace m}e^{-mg} (4.1)
×[i=1m(e𝜹(i)d𝜹(i))]exp(eg𝜹:𝜹𝜹(i) for some ieδ+d𝜹),\displaystyle{}\qquad\times\int\!\left[\prod_{i=1}^{m}\left(e^{\|\bm{\delta}^{(i)}\|}\,\mathrm{d}\bm{\delta}^{(i)}\right)\right]\exp\!\left(-e^{-g}\int_{\bm{\delta}:\bm{\delta}\prec\bm{\delta}^{(i)}\mbox{\scriptsize\ for some\leavevmode\nobreak\ $i$}}e^{\delta_{+}}\,\mathrm{d}\bm{\delta}\right),

where the unlabeled integral is over vectors 𝜹(1),,𝜹(m)\bm{\delta}^{(1)},\ldots,\bm{\delta}^{(m)} satisfying the restrictions (3.16).

Next, multiply both sides of this equality by (Gdg)=exp(eg)egdg\operatorname{\mathbb{P}{}}(G\in\mathrm{d}g)=\exp\!\left(-e^{-g}\right)e^{-g}\,\mathrm{d}g and integrate over gg\in\mathbb{R} to find

𝔼[(𝒦(Δ))r]=m=1r{rm}m!i=1m(e𝜹(i)d𝜹(i))[𝜹:𝜹𝜹(i) for some ieδ+d𝜹]m+1.\operatorname{\mathbb{E}{}}\left[\left(\mathcal{K}(\Delta-)\right)^{r}\right]=\sum_{m=1}^{r}{r\brace m}m!\int\!\frac{\prod_{i=1}^{m}\left(e^{\|\bm{\delta}^{(i)}\|}\,\mathrm{d}\bm{\delta}^{(i)}\right)}{\left[\int_{\bm{\delta}:\bm{\delta}\prec\bm{\delta}^{(i)}\mbox{\scriptsize\ for some\leavevmode\nobreak\ $i$}}e^{\delta_{+}}\,\mathrm{d}\bm{\delta}\right]^{m+1}}.

Changing variables by scaling,

𝔼[(𝒦(Δ))r]=m=1r{rm}m!mdmi=1m(em𝜹(i)d𝜹(i))[𝜹:𝜹m𝜹(i) for some ieδ+d𝜹]m+1,\operatorname{\mathbb{E}{}}\left[\left(\mathcal{K}(\Delta-)\right)^{r}\right]=\sum_{m=1}^{r}{r\brace m}m!\,m^{dm}\int\!\frac{\prod_{i=1}^{m}\left(e^{m\|\bm{\delta}^{(i)}\|}\,\mathrm{d}\bm{\delta}^{(i)}\right)}{\left[\int_{\bm{\delta}:\bm{\delta}\prec m\bm{\delta}^{(i)}\mbox{\scriptsize\ for some\leavevmode\nobreak\ $i$}}e^{\delta_{+}}\,\mathrm{d}\bm{\delta}\right]^{m+1}},

where now the unlabeled integral is over 𝜹(1),,𝜹(m)\bm{\delta}^{(1)},\dots,\bm{\delta}^{(m)} satisfying the restrictions

𝜹(1),,𝜹(m)\bm{\delta}^{(1)},\dots,\bm{\delta}^{(m)} are incomparable, and,
for i=1,,m, we have 𝟎𝜹(i) and 𝜹(i)Δ/m.\displaystyle{}\qquad\quad\mbox{for $i=1,\dots,m$, we have ${\bf 0}\prec\bm{\delta}^{(i)}$ and $\|\bm{\delta}^{(i)}\|\leq\Delta/m$}.

Observe that for each j=1,,mj=1,\dots,m we have

𝜹:𝜹m𝜹(i) for some ieδ+d𝜹𝜹:𝜹m𝜹(j)eδ+d𝜹=em𝜹(j)\int_{\bm{\delta}:\bm{\delta}\prec m\bm{\delta}^{(i)}\mbox{\scriptsize\ for some\leavevmode\nobreak\ $i$}}e^{\delta_{+}}\,\mathrm{d}\bm{\delta}\geq\int_{\bm{\delta}:\bm{\delta}\prec m\bm{\delta}^{(j)}}e^{\delta_{+}}\,\mathrm{d}\bm{\delta}=e^{m\|\bm{\delta}^{(j)}\|}

and hence, taking the geometric mean of the lower bounds,

𝜹:𝜹m𝜹(i) for some ieδ+d𝜹\displaystyle\int_{\bm{\delta}:\bm{\delta}\prec m\bm{\delta}^{(i)}\mbox{\scriptsize\ for some\leavevmode\nobreak\ $i$}}e^{\delta_{+}}\,\mathrm{d}\bm{\delta} j=1me𝜹(j).\displaystyle\geq\prod_{j=1}^{m}e^{\|\bm{\delta}^{(j)}\|}. (4.2)

Thus

𝔼[(𝒦(Δ))r]\displaystyle\operatorname{\mathbb{E}{}}\left[\left(\mathcal{K}(\Delta-)\right)^{r}\right]
m=1r{rm}m!mdmi=1m(e𝜹(i)d𝜹(i))\displaystyle\leq\sum_{m=1}^{r}{r\brace m}\,m!\,m^{dm}\int\!\prod_{i=1}^{m}\left(e^{-\|\bm{\delta}^{(i)}\|}\,\mathrm{d}\bm{\delta}^{(i)}\right)
=m=1r{rm}m!mdm(rm=m;𝐗(i)Δ/m for i=1,,m),\displaystyle=\sum_{m=1}^{r}{r\brace m}\,m!\,m^{dm}\operatorname{\mathbb{P}{}}(r_{m}=m;\,\|\mathbf{X}^{(i)}\|\leq\Delta/m\mbox{\ for $i=1,\dots,m$}), (4.3)

where rnrn(d)r_{n}\equiv r_{n}(d) denotes the number of remaining records at time nn; recall Definition 1.1(b).

We therefore have the bound

𝔼[(𝒦(Δ))r]\displaystyle\operatorname{\mathbb{E}{}}\left[\left(\mathcal{K}(\Delta-)\right)^{r}\right] m=1rm!{rm}mdm(rm=m)\displaystyle\leq\sum_{m=1}^{r}m!{r\brace m}m^{dm}\,\operatorname{\mathbb{P}{}}(r_{m}=m)
m=1rm!{rm}mdmadmm1d1m,\displaystyle\leq\sum_{m=1}^{r}m!{r\brace m}m^{dm}\,a_{d}^{m}\,m^{-\frac{1}{d-1}m},

with ada_{d} as in the statement of the proposition. The inequality

(rm=m)admm1d1m\operatorname{\mathbb{P}{}}(r_{m}=m)\leq a_{d}^{m}\,m^{-\frac{1}{d-1}m} (4.4)

used here is due to Brightwell [2, Theorem 1] [who also gives a lower bound of matching form on (rm=m)\operatorname{\mathbb{P}{}}(r_{m}=m)], answering a question raised by Winkler [7].

Continuing by using the simple upper bound {rm}mr/m!{r\brace m}\leq m^{r}/m!, for any Δ>0\Delta>0 we have

𝔼[(𝒦(Δ))r]\displaystyle\operatorname{\mathbb{E}{}}\left[\left(\mathcal{K}(\Delta-)\right)^{r}\right] m=1rmr+(d1d1)madm\displaystyle\leq\sum_{m=1}^{r}m^{r+(d-\frac{1}{d-1})m}\,a_{d}^{m}
adrr(d+21d1)r<.\displaystyle\leq a_{d}^{r}r^{(d+2-\frac{1}{d-1})r}<\infty. (4.5)

Since 𝒦(Δ)𝒦\mathcal{K}(\Delta-)\uparrow\mathcal{K} as Δ\Delta\uparrow\infty, the proposition follows by the monotone convergence theorem. ∎

Remark 4.2.

Except for first moments [recalling Lemma 3.2(i) and calculating 𝔼𝒦=1\operatorname{\mathbb{E}{}}\mathcal{K}=1 by using the first sentence in the proof of Lemma 3.9 with r=1r=1, integrating out gg with respect to (G){\mathcal{L}}(G), and passing to the limit as Δ\Delta\to\infty] we have not investigated whether the (also finite) moments of KnK_{n} converge to the corresponding moments of 𝒦\mathcal{K}.

Remark 4.3.

(a) When d=2d=2, the logarithmic asymptotics of the bound in Proposition 4.1 do not have the correct coefficient for the lead-order term. Indeed, the moments of 𝒦𝒦(2)\mathcal{K}\equiv\mathcal{K}(2) (see Corollary 6.1) are

𝔼𝒦r=r=1mm!{rm}=12Lir(12)π/2L2rr+(1/2)(eL2)r,\operatorname{\mathbb{E}{}}\mathcal{K}^{r}=\sum_{r=1}^{m}m!{r\brace m}=\frac{1}{2}\operatorname{Li}_{-r}\!\left(\frac{1}{2}\right)\sim\frac{\sqrt{\pi/2}}{\operatorname{L}2}r^{r+(1/2)}(e\operatorname{L}2)^{-r}, (4.6)

where Li\operatorname{Li} denotes polylogarithm.

(b) We do not know, for any d3d\geq 3, whether the logarithmic asymptotics of Proposition 4.1 are correct to lead order. The bound (4.2) is rather crude. However, see Remark 4.7(b).

4.2. Tail asymptotics for (𝒦){\mathcal{L}}(\mathcal{K})

The next two theorems are the main results of this subsection. They give closely (but not perfectly) matching upper and lower bounds on lead-order logarithmic asymptotics for the tail of 𝒦(d)\mathcal{K}(d). We do not presently know for any d3d\geq 3 how to close the gap.

Theorem 4.4.

Fix d2d\geq 2. Then

(𝒦(d)k)exp[Ω(k1/[d+2(d1)1])] as k\operatorname{\mathbb{P}{}}(\mathcal{K}(d)\geq k)\leq\exp\left[-\Omega\!\left(k^{1/[d+2-(d-1)^{-1}]}\right)\right]\mbox{\rm\ \ as $k\to\infty$}

The exponent of kk here can be written as (d1)/(d2+d3)(d-1)/(d^{2}+d-3).

Theorem 4.5.

Fix d2d\geq 2. Then

(𝒦(d)k)exp[O(k1/(d1))] as k.\operatorname{\mathbb{P}{}}(\mathcal{K}(d)\geq k)\geq\exp\left[-O\!\left(k^{1/(d-1)}\right)\right]\mbox{\rm\ \ as $k\to\infty$}.
Proof of Theorem 4.4.

This follows simply from Proposition 4.1, applying Markov’s inequality with an integer-rounding of the optimal choice (in relation to the bound of the proposition)

r=e1(kad)1/(d+21d1),r=e^{-1}\left(\frac{k}{a_{d}}\right)^{1/(d+2-\frac{1}{d-1})},

to wit:

(𝒦(d)k)\displaystyle\operatorname{\mathbb{P}{}}(\mathcal{K}(d)\geq k) kr𝔼𝒦(d)rkr×adrr(d+21d1)r\displaystyle\leq k^{-r}\operatorname{\mathbb{E}{}}\mathcal{K}(d)^{r}\leq k^{-r}\times a_{d}^{r}r^{(d+2-\frac{1}{d-1})r}
=exp[(1+o(1))(d+21d1)1e(kad)1/(d+21d1)].\displaystyle=\exp\left[-(1+o(1))\frac{(d+2-\frac{1}{d-1})^{-1}}{e}\left(\frac{k}{a_{d}}\right)^{1/(d+2-\frac{1}{d-1})}\right]. (4.7)

Remark 4.6.

When d=2d=2, the truth, according to Corollary 6.1, is

(𝒦(2)k)=exp[(L2)k].\operatorname{\mathbb{P}{}}(\mathcal{K}(2)\geq k)=\exp[-(\operatorname{L}2)k].
Proof of Theorem 4.5.

We will establish the stronger assertion that

(𝒦k)exp[(1+o(1))(ck)1/(d1)]\operatorname{\mathbb{P}{}}(\mathcal{K}\geq k)\geq\exp\!\left[-(1+o(1))(ck)^{1/(d-1)}\right] (4.8)

for c=ee1(d1)!c=\frac{e}{e-1}(d-1)! by establishing it for any c>ee1(d1)!c>\frac{e}{e-1}(d-1)!. The idea of the proof is that when G=gG=g is large, 𝒦=𝒦g\mathcal{K}=\mathcal{K}_{g} will be large because 𝒦g𝒦g(g)\mathcal{K}_{g}\geq\mathcal{K}_{g}(g-) and 𝒦g(g)\mathcal{K}_{g}(g-) will be large.

Choose and fix ccd>ee1(d1)!c\equiv c_{d}>\frac{e}{e-1}(d-1)!. Define gk:=(ck)1/(d1)g_{k}:=(ck)^{1/(d-1)}. Then

(𝒦k)\displaystyle\operatorname{\mathbb{P}{}}(\mathcal{K}\geq k) gkgk+1(Gdg)(𝒦gk)\displaystyle\geq\int_{g_{k}}^{g_{k}+1}\operatorname{\mathbb{P}{}}(G\in\mathrm{d}g)\operatorname{\mathbb{P}{}}(\mathcal{K}_{g}\geq k)
gkgk+1(Gdg)(𝒦g(g)k)\displaystyle\geq\int_{g_{k}}^{g_{k}+1}\operatorname{\mathbb{P}{}}(G\in\mathrm{d}g)\operatorname{\mathbb{P}{}}(\mathcal{K}_{g}(g-)\geq k)
=(G(gk,gk+1))\displaystyle=\operatorname{\mathbb{P}{}}(G\in(g_{k},g_{k}+1))
×gkgk+1(GdgG(gk,gk+1))(𝒦g(g)k).\displaystyle{}\qquad\times\int_{g_{k}}^{g_{k}+1}\operatorname{\mathbb{P}{}}(G\in\mathrm{d}g\mid G\in(g_{k},g_{k}+1))\operatorname{\mathbb{P}{}}(\mathcal{K}_{g}(g-)\geq k). (4.9)

For the first factor here, we use (with stated asymptotics as kk\to\infty)

(G(gk,gk+1))\displaystyle\operatorname{\mathbb{P}{}}(G\in(g_{k},g_{k}+1)) =exp[e(gk+1)]exp[egk]\displaystyle=\exp\!\left[-e^{-(g_{k}+1)}\right]-\exp[-e^{-g_{k}}]
=(1+o(1))egk(1+o(1))e(gk+1)\displaystyle=(1+o(1))e^{-g_{k}}-(1+o(1))e^{-(g_{k}+1)}
(1e1)exp[(ck)1/(d1)]\displaystyle\sim(1-e^{-1})\exp\!\left[-(ck)^{1/(d-1)}\right]
=exp{[(ck)1/(d1)+O(1)]}\displaystyle=\exp\!\left\{-\left[(ck)^{1/(d-1)}+O(1)\right]\right\}
=exp[(1+o(1))(ck)1/(d1)].\displaystyle=\exp\!\left[-(1+o(1))(ck)^{1/(d-1)}\right].

We will show that the second factor equals 1o(1)1-o(1). It then follows that

ln(𝒦k)(1+o(1))(ck)1/(d1),-\ln\operatorname{\mathbb{P}{}}(\mathcal{K}\geq k)\leq(1+o(1))(ck)^{1/(d-1)},

as claimed.

It remains to show that

gkgk+1(GdgG(gk,gk+1))(𝒦g(g)k)=1o(1).\int_{g_{k}}^{g_{k}+1}\operatorname{\mathbb{P}{}}(G\in\mathrm{d}g\mid G\in(g_{k},g_{k}+1))\operatorname{\mathbb{P}{}}(\mathcal{K}_{g}(g-)\geq k)=1-o(1).

We will do this by applying Chebyshev’s inequality to the integrand factor (𝒦g(g)k)\operatorname{\mathbb{P}{}}(\mathcal{K}_{g}(g-)\geq k), so to prepare we will obtain a lower bound on 𝔼𝒦g(g)\operatorname{\mathbb{E}{}}\mathcal{K}_{g}(g-) and an upper bound on Var𝒦g(g)\operatorname{Var}\mathcal{K}_{g}(g-).

After some straightforward simplification, it follows from the case r=1r=1 (with Δ=g\Delta=g) of the calculation of 𝔼[(𝒦g(Δ))r]=(3.24)\operatorname{\mathbb{E}{}}[(\mathcal{K}_{g}(\Delta-))^{r}]=\eqref{rth moment} in the proof of Lemma 3.9 and a change of variables (in what follows) from η\eta to v=egeηv=e^{-g}e^{\eta} that, for any 0<ϵ<10<\epsilon<1 and uniformly for g(gk,gk+1)g\in(g_{k},g_{k}+1), we have

𝔼𝒦g(g)\displaystyle\operatorname{\mathbb{E}{}}\mathcal{K}_{g}(g-) =exp(eg)eg0gηd1(d1)!eηexp(egeη)dη\displaystyle=\exp\!\left(e^{-g}\right)e^{-g}\int_{0}^{g}\!\frac{\eta^{d-1}}{(d-1)!}e^{\eta}\exp\!\left(-e^{-g}e^{\eta}\right)\,\mathrm{d}\eta
=exp(eg)gd1(d1)!eg1(1+Lvg)d1evdv\displaystyle=\exp\!\left(e^{-g}\right)\frac{g^{d-1}}{(d-1)!}\int_{e^{-g}}^{1}\left(1+\frac{\operatorname{L}v}{g}\right)^{d-1}e^{-v}\,\mathrm{d}v
exp(eg)gd1(d1)!eϵg1(1ϵ)d1evdv\displaystyle\geq\exp\!\left(e^{-g}\right)\frac{g^{d-1}}{(d-1)!}\int_{e^{-\epsilon g}}^{1}(1-\epsilon)^{d-1}e^{-v}\,\mathrm{d}v
exp[e(gk+1)]gkd1(d1)!(1ϵ)d1[exp(eϵgk)e1]\displaystyle\geq\exp\!\left[e^{-(g_{k}+1)}\right]\frac{g_{k}^{d-1}}{(d-1)!}(1-\epsilon)^{d-1}\left[\exp\!\left(-e^{-\epsilon g_{k}}\right)-e^{-1}\right]
=(1+o(1))(1ϵ)d1c(d1)!(1e1)k.\displaystyle=(1+o(1))(1-\epsilon)^{d-1}\frac{c}{(d-1)!}(1-e^{-1})k.

Thus, uniformly for g(gk,gk+1)g\in(g_{k},g_{k}+1), we have

𝔼𝒦g(g)(1+o(1))c(d1)!(1e1)k.\operatorname{\mathbb{E}{}}\mathcal{K}_{g}(g-)\geq(1+o(1))\frac{c}{(d-1)!}(1-e^{-1})k. (4.10)

Observe that the asymptotic coefficient of kk in (4.10) exceeds 11.

We next turn our attention to the second moment of 𝒦g(g)\mathcal{K}_{g}(g-). From the case r=2r=2 (with Δ=g\Delta=g) of the calculation of 𝔼[(𝒦g(Δ))r]=(3.24)\operatorname{\mathbb{E}{}}[(\mathcal{K}_{g}(\Delta-))^{r}]=\eqref{rth moment} in the proof of Lemma 3.9,

𝔼[(𝒦g(Δ))2]\displaystyle\operatorname{\mathbb{E}{}}\left[\left(\mathcal{K}_{g}(\Delta-)\right)^{2}\right]
=exp(eg)m=12emg\displaystyle=\exp\!\left(e^{-g}\right)\sum_{m=1}^{2}e^{-mg}
×[i=1m(e𝜹(i)d𝜹(i))]exp(eg=1m(1)11i1<i2<<ime𝜹(i1,,il))\displaystyle{}\times\int\!\left[\prod_{i=1}^{m}\left(e^{\|\bm{\delta}^{(i)}\|}\,\mathrm{d}\bm{\delta}^{(i)}\right)\right]\exp\!\left(-e^{-g}\sum_{\ell=1}^{m}(-1)^{\ell-1}\sum_{1\leq i_{1}<i_{2}<\dots<i_{\ell}\leq m}e^{\|\bm{\delta}^{(i_{1},\dots,i_{l})}\|}\right)

where the unlabeled integral is over vectors 𝜹(1),,𝜹(m)\bm{\delta}^{(1)},\dots,\bm{\delta}^{(m)} satisfying the restrictions (3.16) with Δ=g\Delta=g.

Uniformly for g(gk,gk+1)g\in(g_{k},g_{k}+1), the m=1m=1 contribution to this second moment equals

𝔼𝒦g(g)\displaystyle\operatorname{\mathbb{E}{}}\mathcal{K}_{g}(g-) exp(egk)(gk+1)d1(d1)!e(gk+1)1(1+Lvgk)d1evdv\displaystyle\leq\exp\!\left(e^{-g_{k}}\right)\frac{(g_{k}+1)^{d-1}}{(d-1)!}\int_{e^{-(g_{k}+1)}}^{1}\left(1+\frac{\operatorname{L}v}{g_{k}}\right)^{d-1}e^{-v}\,\mathrm{d}v
(1+o(1))(1e1)gkd1(d1)!\displaystyle\leq(1+o(1))(1-e^{-1})\frac{g_{k}^{d-1}}{(d-1)!}
=(1+o(1))c(d1)!(1e1)k.\displaystyle=(1+o(1))\frac{c}{(d-1)!}(1-e^{-1})k. (4.11)

[We remark in passing that the asymptotic bounds (4.10) and (4.11) match.]

The m=2m=2 contribution, call it C2C_{2}, is

exp(eg)e2g\displaystyle\exp\!\left(e^{-g}\right)e^{-2g}
×e𝜹(1)+𝜹(2)exp[eg(e𝜹(1)+e𝜹(1)e𝜹(1,2))]d𝜹(1)d𝜹(2).\displaystyle\times\int\!e^{\|\bm{\delta}^{(1)}\|+\|\bm{\delta}^{(2)}\|}\exp\!\left[-e^{-g}\left(e^{\|\bm{\delta}^{(1)}\|}+e^{\|\bm{\delta}^{(1)}\|}-e^{\|\bm{\delta}^{(1,2)}\|}\right)\right]\,\mathrm{d}\bm{\delta}^{(1)}\,\mathrm{d}\bm{\delta}^{(2)}.

Here we recall that 𝜹(1)\bm{\delta}^{(1)} and 𝜹(2)\bm{\delta}^{(2)} are incomparable. If, for example, 1jd11\leq j\leq d-1 and ϵi(1):=δi(2)δi(1)>0\epsilon^{(1)}_{i}:=\delta^{(2)}_{i}-\delta^{(1)}_{i}>0 for i=1,,ji=1,\dots,j and ϵi(2):=δi(1)δi(2)>0\epsilon^{(2)}_{i}:=\delta^{(1)}_{i}-\delta^{(2)}_{i}>0 for i=j+1,,di=j+1,\dots,d, then, with ϵ:=𝜹(1,2)\bm{\epsilon}:=\bm{\delta}^{(1,2)}, the integrand equals

e2ϵ+ϵ(1)+ϵ(2)exp[egeϵ(eϵ(1)+eϵ(2)1)],e^{2\|\bm{\epsilon}\|+\|\bm{\epsilon}^{(1)}\|+\|\bm{\epsilon}^{(2)}\|}\exp\!\left[-e^{-g}e^{\|\bm{\epsilon}\|}\left(e^{\|\bm{\epsilon}^{(1)}\|}+e^{\|\bm{\epsilon}^{(2)}\|}-1\right)\right],

where ϵ\bm{\epsilon}, ϵ(1)\bm{\epsilon}^{(1)}, and ϵ(2)\bm{\epsilon}^{(2)} are vectors of length dd, jj, and djd-j, respectively. By this reasoning, C2C_{2} equals

exp(eg)e2gj=1d1{(dj)\displaystyle\exp\!\left(e^{-g}\right)e^{-2g}\sum_{j=1}^{d-1}\left\{{d\choose j}\right.
×e2ϵ+ϵ(1)+ϵ(2)exp[egeϵ(eϵ(1)+eϵ(2)1)]dϵ(1)dϵ(2)dϵ},\displaystyle\times\left.\int\!e^{2\|\bm{\epsilon}\|+\|\bm{\epsilon}^{(1)}\|+\|\bm{\epsilon}^{(2)}\|}\exp\!\left[-e^{-g}e^{\|\bm{\epsilon}\|}\left(e^{\|\bm{\epsilon}^{(1)}\|}+e^{\|\bm{\epsilon}^{(2)}\|}-1\right)\right]\,\mathrm{d}\bm{\epsilon}^{(1)}\,\mathrm{d}\bm{\epsilon}^{(2)}\,\mathrm{d}\bm{\epsilon}\right\},

where the integral is over vectors ϵ,ϵ(1),ϵ(2)0\bm{\epsilon},\bm{\epsilon}^{(1)},\bm{\epsilon}^{(2)}\succ 0 of lengths as previously specified, subject to the two restrictions ϵ+ϵ(i)g\|\bm{\epsilon}\|+\|\bm{\epsilon}^{(i)}\|\leq g for i=1,2i=1,2. The integral here reduces to the following three-dimensional integral:

0gηd1(d1)!0gη0gηη1j1(j1)!η2dj1(dj1)!\displaystyle\int_{0}^{g}\!\frac{\eta^{d-1}}{(d-1)!}\int_{0}^{g-\eta}\int_{0}^{g-\eta}\frac{\eta_{1}^{j-1}}{(j-1)!}\frac{\eta_{2}^{d-j-1}}{(d-j-1)!}
×e2η+η1+η2exp[egeη(eη1+eη21)]dη1dη2dη\displaystyle\qquad\qquad\qquad\qquad\qquad\times e^{2\eta+\eta_{1}+\eta_{2}}\exp\!\left[-e^{-g}e^{\eta}\left(e^{\eta_{1}}+e^{\eta_{2}}-1\right)\right]\,\mathrm{d}\eta_{1}\,\mathrm{d}\eta_{2}\,\mathrm{d}\eta
(d2d1j)(d2)!0gηd1(gη)d2(d1)!\displaystyle\leq\frac{{{d-2}\choose{d-1-j}}}{(d-2)!}\int_{0}^{g}\frac{\eta^{d-1}(g-\eta)^{d-2}}{(d-1)!}
×0gη0gηe2η+η1+η2exp[egeη(eη1+eη21)]dη1dη2dη\displaystyle\qquad\qquad\quad\times\int_{0}^{g-\eta}\int_{0}^{g-\eta}e^{2\eta+\eta_{1}+\eta_{2}}\exp\!\left[-e^{-g}e^{\eta}\left(e^{\eta_{1}}+e^{\eta_{2}}-1\right)\right]\,\mathrm{d}\eta_{1}\,\mathrm{d}\eta_{2}\,\mathrm{d}\eta
=(d2d1j)(d2)!0gηd1(gη)d2(d1)!e2ηexp[e(gη)]\displaystyle=\frac{{{d-2}\choose{d-1-j}}}{(d-2)!}\int_{0}^{g}\frac{\eta^{d-1}(g-\eta)^{d-2}}{(d-1)!}e^{2\eta}\exp[e^{-(g-\eta)}]
×{0gηeη1exp[e(gη)eη1]dη1}2dη.\displaystyle\qquad\qquad\qquad\times\left\{\int_{0}^{g-\eta}e^{\eta_{1}}\exp\!\left[-e^{-(g-\eta)}e^{\eta_{1}}\right]\,\mathrm{d}\eta_{1}\right\}^{2}\,\mathrm{d}\eta.

It follows that C2C_{2} is bounded above by exp(eg)\exp\!\left(e^{-g}\right) times

(2d2d1)(d1)!(d2)!e2g0gηd1(gη)d2e2ηexp[e(gη)]\displaystyle\frac{{{2d-2}\choose{d-1}}}{(d-1)!(d-2)!}e^{-2g}\int_{0}^{g}\eta^{d-1}(g-\eta)^{d-2}e^{2\eta}\exp[e^{-(g-\eta)}]
×{0gηeη1exp[e(gη)eη1]dη1}2dη\displaystyle\qquad\qquad\qquad\qquad\quad\times\left\{\int_{0}^{g-\eta}e^{\eta_{1}}\exp\!\left[-e^{-(g-\eta)}e^{\eta_{1}}\right]\,\mathrm{d}\eta_{1}\right\}^{2}\,\mathrm{d}\eta
=(2d2d1)(d1)!(d2)!e2g0gηd1(gη)d2e2ηexp[e(gη)]\displaystyle=\frac{{{2d-2}\choose{d-1}}}{(d-1)!(d-2)!}e^{-2g}\int_{0}^{g}\eta^{d-1}(g-\eta)^{d-2}e^{2\eta}\exp[e^{-(g-\eta)}]
×{egη(exp[e(gη)]e1)}2dη\displaystyle\qquad\qquad\qquad\qquad\qquad\qquad\times\left\{e^{g-\eta}\left(\exp\!\left[-e^{-(g-\eta)}\right]-e^{-1}\right)\right\}^{2}\,\mathrm{d}\eta
(2d2d1)(1e1)2(d1)!(d2)!0gηd1(gη)d2exp[e(gη)]dη\displaystyle\leq\frac{{{2d-2}\choose{d-1}}(1-e^{-1})^{2}}{(d-1)!(d-2)!}\int_{0}^{g}\eta^{d-1}(g-\eta)^{d-2}\exp[e^{-(g-\eta)}]\,\mathrm{d}\eta
=(2d2d1)(1e1)2(d1)!(d2)!g2d201td1(1t)d2exp[e(1t)g]dt.\displaystyle=\frac{{{2d-2}\choose{d-1}}(1-e^{-1})^{2}}{(d-1)!(d-2)!}g^{2d-2}\int_{0}^{1}\!t^{d-1}(1-t)^{d-2}\exp[e^{-(1-t)g}]\,\mathrm{d}t.

By the dominated convergence theorem, as gg\to\infty we have

C2C2(g)(1+o(1))αdg2d2,C_{2}\equiv C_{2}(g)\leq(1+o(1))\alpha_{d}g^{2d-2},

where

αd=(1e1)2(2d2d1)01td1(1t)d2dt(d1)!(d2)!=[1e1(d1)!]2.\alpha_{d}=(1-e^{-1})^{2}{{2d-2}\choose{d-1}}\frac{\int_{0}^{1}\!t^{d-1}(1-t)^{d-2}\,\mathrm{d}t}{{(d-1)!(d-2)!}}=\left[\frac{1-e^{-1}}{(d-1)!}\right]^{2}.

In particular, uniformly for g(gk,gk+1)g\in(g_{k},g_{k}+1), we have

C2(g)(1+o(1))[(1e1)gkd1(d1)!]2[c(d1)!(1e1)k]2.C_{2}(g)\leq(1+o(1))\left[(1-e^{-1})\frac{g_{k}^{d-1}}{(d-1)!}\right]^{2}\sim\left[\frac{c}{(d-1)!}(1-e^{-1})k\right]^{2}. (4.12)

We conclude from (4.11), (4.12), and (4.10) that

Var𝒦g(g)=o(k2),\operatorname{Var}\mathcal{K}_{g}(g-)=o(k^{2}),

uniformly for g(gk,gk+1)g\in(g_{k},g_{k}+1).

By Chebyshev’s inequality, uniformly for g(gk,gk+1)g\in(g_{k},g_{k}+1) we have

(𝒦g(g)k)1Var𝒦g(g)[𝔼𝒦g(g)k]2=1o(k2)Θ(k2)=1o(1),\operatorname{\mathbb{P}{}}(\mathcal{K}_{g}(g-)\geq k)\geq 1-\frac{\operatorname{Var}\mathcal{K}_{g}(g-)}{\left[\operatorname{\mathbb{E}{}}\mathcal{K}_{g}(g-)-k\right]^{2}}=1-\frac{o(k^{2})}{\Theta(k^{2})}=1-o(1),

as was to be shown. ∎

Remark 4.7.

(a) When d=2d=2, the lower bound (4.8) with c=ee1(d1)!c=\frac{e}{e-1}(d-1)! gives

(𝒦k)exp[(1+o(1))ee1k].\operatorname{\mathbb{P}{}}(\mathcal{K}\geq k)\geq\exp\!\left[-(1+o(1))\frac{e}{e-1}k\right].

The logarithmic asymptotics are of the correct order (namely, linear), but the coefficient e/(e1)e/(e-1) is approximately 1.5821.582, which is roughly twice as big as the correct coefficient (recall Remark 4.6) ln20.693\ln 2\doteq 0.693.

(b) Theorem 4.5 can be used to give a lower bound on the moments of 𝒦\mathcal{K} by starting with the lower bound

𝔼𝒦rkr(𝒦k)\operatorname{\mathbb{E}{}}\mathcal{K}^{r}\geq k^{r}\operatorname{\mathbb{P}{}}(\mathcal{K}\geq k)

and choosing kk(r)k\equiv k(r) judiciously. The result for fixed dd, in short, is that

𝔼𝒦rexp[(1+o(1))(d1)rLr] as r.\operatorname{\mathbb{E}{}}\mathcal{K}^{r}\geq\exp[(1+o(1))(d-1)r\operatorname{L}r]\mbox{\ \ as $r\to\infty$}. (4.13)

The lead-order terms of Proposition 4.1 and (4.13) for the logarithmic asymptotics are of the same order, but the coefficients [d+2(d1)1d+2-(d-1)^{-1} and d1d-1, respectively] don’t quite match.

Remark 4.8.

It may be of some interest to study the distribution of 𝒦(Δ)\mathcal{K}(\Delta-) for fixed Δ<\Delta<\infty, but it is then simpler for measuring the distance of a killed record from the new record to switch from 1\ell^{1}-distance to \ell^{\infty} distance. We call the analogue of 𝒦(Δ)\mathcal{K}(\Delta-) for the latter distance 𝒦~(Δ)\widetilde{\mathcal{K}}(\Delta-).

The goal of this remark is to show that, for fixed dd and Δ\Delta, as kk\to\infty we have ln(𝒦~(Δ)k)=Θ(klnk)-\ln\operatorname{\mathbb{P}{}}(\widetilde{\mathcal{K}}(\Delta-)\geq k)=\Theta(k\ln k) (in stark contrast to Theorem 4.5); more precisely, we show (a) that

(𝒦~(Δ)k)exp[ln(11β)ln(21β)(d1)1klnk+O(k)]\operatorname{\mathbb{P}{}}(\widetilde{\mathcal{K}}(\Delta-)\geq k)\leq\exp\!\left[-\frac{\ln\!\left(\frac{1}{1-\beta}\right)}{\ln\!\left(\frac{2}{1-\beta}\right)}(d-1)^{-1}k\ln k+O(k)\right] (4.14)

with ββd,Δ:=[(eΔ1)d+1]1(0,1)\beta\equiv\beta_{d,\Delta}:=[(e^{\Delta}-1)^{d}+1]^{-1}\in(0,1) [so that the increasing function ln(11β)/ln(21β)\ln\!\left(\frac{1}{1-\beta}\right)/\ln\!\left(\frac{2}{1-\beta}\right) of β\beta belongs to (0,1)(0,1) as well] and (b) that

(𝒦~(Δ)k)exp[(d1)1klnk+O(k)]\operatorname{\mathbb{P}{}}(\widetilde{\mathcal{K}}(\Delta-)\geq k)\geq\exp\!\left[-(d-1)^{-1}k\ln k+O(k)\right] (4.15)

(a) Here is a sketch of the proof of (4.14). In order to have 𝒦~(Δ)k\widetilde{\mathcal{K}}(\Delta-)\geq k, the Poisson process NgN_{g} in the proof of Lemma 3.8 must have nn points x\prec x for some nkn\geq k, and at least kk of them must be maxima among these nn points. Integrating over gg, we find

(𝒦~(Δ)k)βn=k(1β)n(rnk)\operatorname{\mathbb{P}{}}(\widetilde{\mathcal{K}}(\Delta-)\geq k)\leq\beta\sum_{n=k}^{\infty}(1-\beta)^{n}\operatorname{\mathbb{P}{}}(r_{n}\geq k)

Over the event {rnk}\{r_{n}\geqslant k\} there must be some kk-tuple of incomparable observations; thus, by finite subadditivity,

(rnk)(nk)(rk=k)2n(rk=k).\operatorname{\mathbb{P}{}}(r_{n}\geq k)\leq{n\choose k}\operatorname{\mathbb{P}{}}(r_{k}=k)\leq 2^{n}\operatorname{\mathbb{P}{}}(r_{k}=k).

Recall from (4.4) that

(rk=k)adkk1d1k.\operatorname{\mathbb{P}{}}(r_{k}=k)\leq a_{d}^{k}k^{-\frac{1}{d-1}k}.

Thus for any k0k0(k)kk_{0}\equiv k_{0}(k)\geq k we have

(𝒦~(Δ)k)\displaystyle\operatorname{\mathbb{P}{}}(\widetilde{\mathcal{K}}(\Delta-)\geq k) βn=kk012nadkk1d1k+βn=k0(1β)n\displaystyle\leq\beta\sum_{n=k}^{k_{0}-1}2^{n}a_{d}^{k}k^{-\frac{1}{d-1}k}+\beta\sum_{n=k_{0}}^{\infty}(1-\beta)^{n}
β2k0adkk1d1k+(1β)k0.\displaystyle\leq\beta 2^{k_{0}}a_{d}^{k}k^{-\frac{1}{d-1}k}+(1-\beta)^{k_{0}}.

A nearly optimal choice of k0k_{0} here is to round [ln(21β)]1(d1)1klnk\left[\ln\!\left(\frac{2}{1-\beta}\right)\right]^{-1}(d-1)^{-1}k\ln k to an integer, and this yields (4.14).

(b) To prove (4.15), we start (in similar fashion as the proof of Theorem 4.5) with a study of 𝒦~g(Δ)\widetilde{\mathcal{K}}_{g}(\Delta-). Using the Poisson-process notation NgN_{g} as we did in the proof of Lemma 3.8, observe that

(𝒦~g(Δ)k)n=k(Ng(A)=n)(rnk)(Ng(B)=0),\operatorname{\mathbb{P}{}}(\widetilde{\mathcal{K}}_{g}(\Delta-)\geq k)\geq\sum_{n=k}^{\infty}\operatorname{\mathbb{P}{}}(N_{g}(A)=n)\operatorname{\mathbb{P}{}}(r_{n}\geq k)\operatorname{\mathbb{P}{}}(N_{g}(B)=0),

where, with 𝐱:=(g/d,,g/d)\mathbf{x}:=(g/d,\dots,g/d) and 𝚫:=(Δ,,Δ)\bm{\Delta}:=(\Delta,\dots,\Delta), we set

A:={𝐳:𝐱𝚫𝐳𝐱},B:={𝐳:𝐱𝚫𝐳 and 𝐱𝐳}.A:=\{\mathbf{z}:\,\mathbf{x}-\bm{\Delta}\prec\mathbf{z}\prec\mathbf{x}\},\qquad B:=\{\mathbf{z}:\mathbf{x}-\bm{\Delta}\prec\mathbf{z}\mbox{\ and\ }\mathbf{x}\not\prec\mathbf{z}\}.

The random variables Ng(A)N_{g}(A) and Ng(B)N_{g}(B) are each Poisson distributed with respective means

λ=eg(eΔ1)d,μ=eg(edΔ1).\lambda=e^{-g}(e^{\Delta}-1)^{d},\qquad\mu=e^{-g}(e^{d\Delta}-1).

Since the number rnr_{n} of remaining records in dimension dd has the same distribution as the total number RnR_{n} of records set through time nn in dimension d1d-1, it follows that rnr_{n} increases stochastically in nn. Thus

(𝒦~g(Δ)k)(Ng(A)k)(rk=k)eμ.\operatorname{\mathbb{P}{}}(\widetilde{\mathcal{K}}_{g}(\Delta-)\geq k)\geq\operatorname{\mathbb{P}{}}(N_{g}(A)\geq k)\operatorname{\mathbb{P}{}}(r_{k}=k)e^{-\mu}.

If ggk:=Lk+cg\leq g_{k}:=-\operatorname{L}k+c with <ccd,Δ<dL(eΔ1)-\infty<c\equiv c_{d,\Delta}<d\operatorname{L}(e^{\Delta}-1), then Chebyshev’s inequality gives (uniformly for such gg) the result

(Ng(A)k)=1o(1)\operatorname{\mathbb{P}{}}(N_{g}(A)\geq k)=1-o(1)

as kk\to\infty. If ggk+1g\geq g_{k+1}, then

eμexp[ec(edΔ1)(k+1)]exp[(eΔ1)1(edΔ1)(k+1)].e^{-\mu}\geq\exp[-e^{-c}(e^{d\Delta}-1)(k+1)]\geq\exp[-(e^{\Delta}-1)^{-1}(e^{d\Delta}-1)(k+1)].

Further, according to [2],

(rk=k)=exp[(d1)1kLk+O(k)]\operatorname{\mathbb{P}{}}(r_{k}=k)=\exp\!\left[-(d-1)^{-1}k\operatorname{L}k+O(k)\right]

as kk\to\infty. To summarize our treatment thus far, uniformly for gk+1ggkg_{k+1}\leq g\leq g_{k} we have

(𝒦~g(Δ)k)exp[(d1)1kLk+O(k)].\operatorname{\mathbb{P}{}}(\widetilde{\mathcal{K}}_{g}(\Delta-)\geq k)\geq\exp\!\left[-(d-1)^{-1}k\operatorname{L}k+O(k)\right]. (4.16)

But if GG is distributed standard Gumbel, we also have

(gk+1Ggk)\displaystyle\operatorname{\mathbb{P}{}}(g_{k+1}\leq G\leq g_{k}) =eeckeec(k+1)=(1eec)eeck\displaystyle=e^{-e^{-c}k}-e^{-e^{-c}(k+1)}=(1-e^{-e^{-c}})e^{-e^{-c}k}
=exp[O(k)].\displaystyle=\exp[-O(k)]. (4.17)

The assertion (4.15) follows from (4.17) and (4.16).

5. Asymptotics of (𝒦(d)){\mathcal{L}}(\mathcal{K}(d)) as dd\to\infty

In this section we prove that 𝒦(d)\mathcal{K}(d) converges in probability (equivalently, in distribution) to 0 as the dimension dd tends to infinity by establishing upper (Subsection 5.1) and lower (Subsection 5.2) bounds, each decaying exponentially to 0, on (𝒦(d)1)\operatorname{\mathbb{P}{}}(\mathcal{K}(d)\geq 1).

5.1. Upper bound

Here is the main result of this subsection, showing that the decay of (𝒦(d)1)\operatorname{\mathbb{P}{}}(\mathcal{K}(d)\geq 1) to 0 is at least exponentially rapid.

Theorem 5.1.

As dd\to\infty we have

(𝒦(d)1)(1+o(1))(0.623)d.\operatorname{\mathbb{P}{}}(\mathcal{K}(d)\geq 1)\leq(1+o(1))\,(0.623)^{d}.
Proof.

Passing to the limit as Δ\Delta\to\infty from (4.1) with r=1r=1, recall that for each gg\in\mathbb{R} we have

𝔼𝒦g\displaystyle\operatorname{\mathbb{E}{}}\mathcal{K}_{g} =exp(eg)eg𝜹(1)𝟎e𝜹(1)exp(ege𝜹(1))d𝜹(1)\displaystyle=\exp\!\left(e^{-g}\right)e^{-g}\int_{\bm{\delta}^{(1)}\succ{\bf 0}}\!e^{\|\bm{\delta}^{(1)}\|}\exp\!\left(-e^{-g}e^{\|\bm{\delta}^{(1)}\|}\right)\,\mathrm{d}\bm{\delta}^{(1)}
=exp(eg)0ηd1(d1)!e(gη)exp[e(gη)]dη.\displaystyle=\exp\!\left(e^{-g}\right)\int_{0}^{\infty}\!\frac{\eta^{d-1}}{(d-1)!}e^{-(g-\eta)}\exp\!\left[-e^{-(g-\eta)}\right]\,\mathrm{d}\eta.

Thus for any g>0g>0 we have

(𝒦(d)1)(|G|>g)+gge2γId(γ)dγ\operatorname{\mathbb{P}{}}(\mathcal{K}(d)\geq 1)\leq\operatorname{\mathbb{P}{}}(|G|>g)+\int_{-g}^{g}\!e^{-2\gamma}I_{d}(\gamma)\,\mathrm{d}\gamma (5.1)

where the integral Id(γ)I_{d}(\gamma) is defined, and can be bounded for any c>0c>0, as follows:

Id(γ)\displaystyle I_{d}(\gamma) :=0ηd1(d1)!eηexp[e(γη)]dη\displaystyle:=\int_{0}^{\infty}\!\frac{\eta^{d-1}}{(d-1)!}e^{\eta}\exp\!\left[-e^{-(\gamma-\eta)}\right]\,\mathrm{d}\eta
c(d1)0e(c+1)ηexp[e(γη)]dη\displaystyle\leq c^{-(d-1)}\int_{0}^{\infty}\!e^{(c+1)\eta}\exp\!\left[-e^{-(\gamma-\eta)}\right]\,\mathrm{d}\eta
c(d1)e(c+1)ηexp[e(γη)]dη\displaystyle\leq c^{-(d-1)}\int_{-\infty}^{\infty}\!e^{(c+1)\eta}\exp\!\left[-e^{-(\gamma-\eta)}\right]\,\mathrm{d}\eta
=c(d1)Γ(c+1)exp[(c+1)γ].\displaystyle=c^{-(d-1)}\Gamma(c+1)\exp[(c+1)\gamma].

Returning to (5.1), for c>1c>1 we find

(𝒦(d)1)\displaystyle\operatorname{\mathbb{P}{}}(\mathcal{K}(d)\geq 1) (|G|>g)+c(d1)Γ(c+1)gge(c1)γdγ\displaystyle\leq\operatorname{\mathbb{P}{}}(|G|>g)+c^{-(d-1)}\Gamma(c+1)\int_{-g}^{g}\!e^{(c-1)\gamma}\,\mathrm{d}\gamma
(|G|>g)+c(d1)Γ(c+1)(c1)1e(c1)g.\displaystyle\leq\operatorname{\mathbb{P}{}}(|G|>g)+c^{-(d-1)}\Gamma(c+1)(c-1)^{-1}e^{(c-1)g}.

As gg\to\infty, this last bound has the asymptotics

(1+o(1))eg+c(d1)Γ(c+1)(c1)1e(c1)g.(1+o(1))e^{-g}+c^{-(d-1)}\Gamma(c+1)(c-1)^{-1}e^{(c-1)g}.

Thus, to obtain an approximately optimal bound we choose

g=c1L[cd1Γ(c+1)].g=c^{-1}\operatorname{L}\left[\frac{c^{d-1}}{\Gamma(c+1)}\right].

This choice gives the following asymptotics as dd\to\infty:

(𝒦(d)1)(1+o(1))cc1[cd1Γ(c+1)]c1.\operatorname{\mathbb{P}{}}(\mathcal{K}(d)\geq 1)\leq(1+o(1))\frac{c}{c-1}\left[\frac{c^{d-1}}{\Gamma(c+1)}\right]^{-c^{-1}}.

The optimal choice of cc for large dd minimizes cc1c^{-c^{-1}}, leading to c=ec=e and

(𝒦(d)1)(1+o(1))(0.623)d,\operatorname{\mathbb{P}{}}(\mathcal{K}(d)\geq 1)\leq(1+o(1))\,(0.623)^{d},

as claimed, since exp[e1]<0.623\exp[-e^{-1}]<0.623. ∎

5.2. Lower bound

Here is the main result of this subsection, showing that the decay of (𝒦(d)1)\operatorname{\mathbb{P}{}}(\mathcal{K}(d)\geq 1) to 0 is at most exponentially rapid.

Theorem 5.2.

For every d1d\geq 1 we have

(𝒦(d)1)4d.\operatorname{\mathbb{P}{}}(\mathcal{K}(d)\geq 1)\geq 4^{-d}.
Proof.

We use Poisson process notation as in the proof of Lemma 3.8. Let 𝟏:=(1,,1)d{\bf 1}:=(1,\dots,1)\in\mathbb{R}^{d}. Given gg\in\mathbb{R} and c>0c>0, let 𝐱(g)=gd𝟏\mathbf{x}(g)=\frac{g}{d}{\bf 1} and

Sc:={𝐳d:𝐳𝐱c𝟏,𝐳𝐱},S_{c}:=\{\mathbf{z}\in\mathbb{R}^{d}:\mathbf{z}\succ\mathbf{x}-c{\bf 1},\,\mathbf{z}\not\succ\mathbf{x}\},

and consider the subset

Sc:={𝐳d:𝐱c𝟏𝐳𝐱}S^{\prime}_{c}:=\{\mathbf{z}\in\mathbb{R}^{d}:\mathbf{x}-c{\bf 1}\prec\mathbf{z}\prec\mathbf{x}\}

of ScS_{c}. Then

(𝒦g1)\displaystyle\operatorname{\mathbb{P}{}}(\mathcal{K}_{g}\geq 1) (Ng(Sc)=1=Ng(Sc))\displaystyle\geq\operatorname{\mathbb{P}{}}(N_{g}(S^{\prime}_{c})=1=N_{g}(S_{c}))
=(Ng(Sc)=1,Ng(ScSc)=0)\displaystyle=\operatorname{\mathbb{P}{}}(N_{g}(S^{\prime}_{c})=1,\,N_{g}(S_{c}-S^{\prime}_{c})=0)
=eλλ11!×exp(𝐳ScScez+d𝐳)\displaystyle=e^{-\lambda}\frac{\lambda^{1}}{1!}\times\exp\!\left(-\int_{\mathbf{z}\in S_{c}-S^{\prime}_{c}}e^{-z_{+}}\,\mathrm{d}\mathbf{z}\right) (5.2)

with

λ=𝐳Scez+d𝐳=i=1d{exp[(xi(g)c)]exp[xi(g)]}=eg(ec1)d.\lambda=\int_{\mathbf{z}\in S^{\prime}_{c}}e^{-z_{+}}\,\mathrm{d}\mathbf{z}=\prod_{i=1}^{d}\left\{\exp\left[-(x_{i}(g)-c)\right]-\exp\left[-x_{i}(g)\right]\right\}=e^{-g}\left(e^{c}-1\right)^{d}.

The integral in (5.2) equals the difference of the integrals over ScS_{c} and over ScS^{\prime}_{c}, namely,

𝐳𝐱c𝟏ez+d𝐳𝐳𝐱ez+d𝐳λ=e(gcd)egλ.\int_{\mathbf{z}\succ\mathbf{x}-c{\bf 1}}e^{-z_{+}}\,\mathrm{d}\mathbf{z}-\int_{\mathbf{z}\succ\mathbf{x}}e^{-z_{+}}\,\mathrm{d}\mathbf{z}-\lambda=e^{-(g-cd)}-e^{-g}-\lambda.

So

(𝒦g1)(ec1)degexp[eg(ecd1)]\operatorname{\mathbb{P}{}}(\mathcal{K}_{g}\geq 1)\geq\left(e^{c}-1\right)^{d}e^{-g}\exp\!\left[-e^{-g}(e^{cd}-1)\right]

and hence

(𝒦(d)1)\displaystyle\operatorname{\mathbb{P}{}}(\mathcal{K}(d)\geq 1) (ec1)de2gexp[egecd]dg\displaystyle\geq\left(e^{c}-1\right)^{d}\int_{-\infty}^{\infty}e^{-2g}\exp\!\left[-e^{-g}e^{cd}\right]\,\mathrm{d}g
=(ec1)de2cd=(ec1e2c)d.\displaystyle=\left(e^{c}-1\right)^{d}e^{-2cd}=\left(\frac{e^{c}-1}{e^{2c}}\right)^{d}.

The optimal choice of cc is then c=ln2c=\ln 2, yielding the claimed result. ∎

Remark 5.3.

In the same spirit as their Conjecture 2.3 and the discussion following it, the authors of [4] might have put forth the closely related conjecture that (Kn=0Kn0)\operatorname{\mathbb{P}{}}(K_{n}=0\!\mid\!K_{n}\geq 0) has a limit qdq_{d} and that qd1q_{d}\to 1 as dd\to\infty, with the additional suggestion that perhaps qd=1d1q_{d}=1-d^{-1} for every d2d\geq 2. In light of Theorems 1.3 and 5.15.2 we see that the related conjecture is true, but the additional suggestion is not.

6. Dimension d=2d=2

Here is the main result of this section.

Corollary 6.1.

Adopt the setting and notation of Theorem 1.3 with d=2d=2. Then

(i) For each gg\in\mathbb{R}, the law of 𝒦g\mathcal{K}_{g} is the mixture 𝔼Poisson(Λg)\operatorname{\mathbb{E}{}}\mbox{\rm Poisson}(\Lambda_{g}), where (Λg){\mathcal{L}}(\Lambda_{g}) is the conditional distribution of gGg-G given gG>0g-G>0 with GG\sim standard Gumbel.

(ii) 𝒦=𝒦(2)\mathcal{K}=\mathcal{K}(2) has the same distribution as 𝒢1\mathcal{G}-1, where 𝒢\mathcal{G} (with support {1,2,}\{1,2,\dots\}) has the Geometric(1/2)(1/2) distribution.

Proof.

We first show how (ii) follows from (i). If (i) holds, then the law of 𝒦\mathcal{K} is the mixture 𝔼Poisson(Λ)\operatorname{\mathbb{E}{}}\mbox{\rm Poisson}(\Lambda), where (Λ){\mathcal{L}}(\Lambda) is the mixture 𝔼(ΛG)\operatorname{\mathbb{E}{}}{\mathcal{L}}(\Lambda_{G}) with GG\sim standard Gumbel. Observe that the density of Λg\Lambda_{g} is

λ𝟏(λ>0)eλgexp[eg(eλ1)],\lambda\mapsto{\bf 1}(\lambda>0)e^{\lambda-g}\exp\!\left[-e^{-g}(e^{\lambda}-1)\right], (6.1)

so the density of Λ\Lambda is

λeλ2gexp[eλg]dg=eλ;\lambda\mapsto\int_{-\infty}^{\infty}e^{\lambda-2g}\exp\!\left[-e^{\lambda-g}\right]\,\mathrm{d}g=e^{-\lambda};

that is, Λ\Lambda has the Exponential(1)(1) distribution. But then (𝒦){\mathcal{L}}(\mathcal{K}) is the law of 𝒢1\mathcal{G}-1, as follows either computationally: for k=0,1,k=0,1,\dots we have

(𝒦=k)=0ewwkk!ewdw=2(k+1);\operatorname{\mathbb{P}{}}(\mathcal{K}=k)=\int_{0}^{\infty}\!e^{-w}\frac{w^{k}}{k!}e^{-w}\,\mathrm{d}w=2^{-(k+1)};

or by the probabilistic argument that 𝒦\mathcal{K} has the same distribution as the number of arrivals in one Poisson counting process with unit rate prior to the (Exponentially distributed) epoch of first arrival in an independent copy of the process, which (using symmetry of the two processes and the memoryless property of the Exponential distribution) has the same distribution as 𝒢1\mathcal{G}-1.

We now proceed to prove (i). It is easy to see that, with probability 11, the Poisson point process (call it NgN_{g}) described in Theorem 1.3 will have an infinite number of maxima, but with no accumulation points in either of the coordinates, and a finite number of maxima dominated by 𝐱=(x,y):=(g/2,g/2)\mathbf{x}=(x,y):=(g/2,g/2). [It is worth noting that the argument to follow is unchanged if 𝐱\mathbf{x} is changed to any other point (x,y)(x,y) with x+y=gx+y=g.] Thus, over the event {𝒦g=k}\{\mathcal{K}_{g}=k\} we can list the locations of the maxima 𝐗i=(Xi,Yi)\mathbf{X}_{i}=(X_{i},Y_{i}), in order from northwest to southeast (i.e., in increasing order of first coordinate and decreasing order of second coordinate), as ,𝐗0,𝐗1,,𝐗k,𝐗k+1,\ldots,\mathbf{X}_{0},\mathbf{X}_{1},\ldots,\mathbf{X}_{k},\mathbf{X}_{k+1},\ldots, where 𝐗1,,𝐗k\mathbf{X}_{1},\ldots,\mathbf{X}_{k} are the maxima dominated by 𝐱\mathbf{x}. Given

<x0<x1<<xk<x<xk+1<\displaystyle\hskip 46.97505pt-\infty<x_{0}<x_{1}<\cdots<x_{k}<x<x_{k+1}<\infty (6.2)
and >y0>y>y1>>yk>yk+1>,\displaystyle\mbox{and\ }\infty>y_{0}>y>y_{1}>\cdots>y_{k}>y_{k+1}>-\infty, (6.3)

let SS denote the following disjoint union of rectangular regions:

S=i=1k[(xi1,xi)×(yi,)][(xk,x)×(yk+1,)][(x,)×(yk+1,y)].S=\bigcup_{i=1}^{k}[(x_{i-1},x_{i})\times(y_{i},\infty)]\,\cup\,[(x_{k},x)\times(y_{k+1},\infty)]\,\cup\,[(x,\infty)\times(y_{k+1},y)].

Then (by the same sort of reasoning as in [3, Proposition 3.1]), for k=0,1,k=0,1,\ldots and 𝐱0,,𝐱k\mathbf{x}_{0},\dots,\mathbf{x}_{k} satisfying (6.2)–(6.3), and introducing abbreviations

jk:=i=jk(exi1exi)eyi,k:=1k,\sum_{j}^{k}:=\sum_{i=j}^{k}(e^{-x_{i-1}}-e^{-x_{i}})e^{-y_{i}},\qquad\sum^{k}:=\sum_{1}^{k},

we have

(𝒦g=k;𝐗id𝐱i for i=0,,k+1)\displaystyle\operatorname{\mathbb{P}{}}(\mathcal{K}_{g}=k;\,\mathbf{X}_{i}\in\mathrm{d}\mathbf{x}_{i}\mbox{\ for $i=0,\ldots,k+1$})
=(Ng(d𝐱i)=1 for i=0,,k+1;Ng(S)=0)\displaystyle=\operatorname{\mathbb{P}{}}(N_{g}(\mathrm{d}\mathbf{x}_{i})=1\mbox{\ for $i=0,\dots,k+1$};N_{g}(S)=0)
=[i=0k+1(e(xi+yi)d𝐱i)]×exp[𝐳Sez+dz]\displaystyle=\left[\prod_{i=0}^{k+1}(e^{-(x_{i}+y_{i})}\,\mathrm{d}\mathbf{x}_{i})\right]\times\exp\!\left[-\int_{\mathbf{z}\in S}\!e^{-z_{+}}\,\mathrm{d}z\right]
=exp[i=0k+1(xi+yi)]exp{eg[k+e(xk+yk+1)]}d𝐱0d𝐱k+1.\displaystyle=\exp\!\left[-\sum_{i=0}^{k+1}(x_{i}+y_{i})\right]\,\exp\!\left\{e^{-g}-\left[\sum^{k}{}+e^{-(x_{k}+y_{k+1})}\right]\right\}\,\mathrm{d}\mathbf{x}_{0}\cdots\,\mathrm{d}\mathbf{x}_{k+1}. (6.4)

To calculate (𝒦g=k)\operatorname{\mathbb{P}{}}(\mathcal{K}_{g}=k), we need to integrate this last expression over all 𝐱0,,𝐱k+1\mathbf{x}_{0},\ldots,\mathbf{x}_{k+1} satisfying (6.2)–(6.3).

Since xk+1x_{k+1} and y0y_{0} appear only in the first of the two factors in (6.4), we integrate them out first to obtain

egexp(i=0kxi)exp(i=1k+1yi)\displaystyle e^{-g}\int\exp\!\left(-\sum_{i=0}^{k}x_{i}\right)\exp\!\left(-\sum_{i=1}^{k+1}y_{i}\right) (6.5)
×exp{eg[k+e(xk+yk+1)]}dx0d𝐱1d𝐱kdyk+1,\displaystyle{}\qquad\times\exp\!\left\{e^{-g}-\left[\sum^{k}{}+e^{-(x_{k}+y_{k+1})}\right]\right\}\,\mathrm{d}x_{0}\,\mathrm{d}\mathbf{x}_{1}\cdots\,\mathrm{d}\mathbf{x}_{k}\,\mathrm{d}y_{k+1},

where the integral is over all x0,𝐱1,𝐱k,yk+1x_{0},\mathbf{x}_{1},\ldots\mathbf{x}_{k},y_{k+1} satisfying

<x0<x1<<xk<x\displaystyle\hskip 36.135pt-\infty<x_{0}<x_{1}<\cdots<x_{k}<x
and y>y1>>yk>yk+1>.\displaystyle\mbox{and\ }y>y_{1}>\cdots>y_{k}>y_{k+1}>-\infty.

Next we integrate out x0x_{0}. For this we use the calculation

x1ex0exp(ey1ex0)dx0=ey1exp(ex1ey1).\int_{-\infty}^{x_{1}}e^{-x_{0}}\exp\!\left(-e^{-y_{1}}e^{-x_{0}}\right)\,\mathrm{d}x_{0}=e^{y_{1}}\exp\!\left(-e^{-x_{1}}e^{-y_{1}}\right).

So when we integrate out x0x_{0} we get

egexp(i=1kxi)exp(i=2k+1yi)\displaystyle e^{-g}\int\exp\!\left(-\sum_{i=1}^{k}x_{i}\right)\exp\!\left(-\sum_{i=2}^{k+1}y_{i}\right)
×exp{eg[2k+e(xk+yk+1)]}d𝐱1d𝐱kdyk+1,\displaystyle{}\qquad\times\exp\!\left\{e^{-g}-\left[\sum_{2}^{k}{}+e^{-(x_{k}+y_{k+1})}\right]\right\}\,\mathrm{d}\mathbf{x}_{1}\cdots\,\mathrm{d}\mathbf{x}_{k}\,\mathrm{d}y_{k+1},

where the integral is over all 𝐱1,𝐱k,yk+1\mathbf{x}_{1},\ldots\mathbf{x}_{k},y_{k+1} satisfying

<x1<<xk<x\displaystyle\hskip 36.135pt-\infty<x_{1}<\cdots<x_{k}<x
and y>y1>>yk>yk+1>.\displaystyle\mbox{and\ }y>y_{1}>\cdots>y_{k}>y_{k+1}>-\infty.

Continuing in this fashion, after integrating out x1,,xkx_{1},\dots,x_{k} we get

egexp{ege(x+yk+1)}dy1dykdyk+1,e^{-g}\int\exp\!\left\{e^{-g}-e^{-(x+y_{k+1})}\right\}\,\mathrm{d}y_{1}\cdots\,\mathrm{d}y_{k}\,\mathrm{d}y_{k+1},

where the integral is over all y1,,yk+1y_{1},\dots,y_{k+1} satisfying

y>y1>>yk>yk+1>.y>y_{1}>\cdots>y_{k}>y_{k+1}>-\infty.

Next we integrate out y1,,yky_{1},\dots,y_{k} and change the remaining variable name from yky_{k} to η\eta to find

(𝒦g=k)\displaystyle\operatorname{\mathbb{P}{}}(\mathcal{K}_{g}=k) =egk!y(yη)kexp{ege(x+η)}dη\displaystyle=\frac{e^{-g}}{k!}\int_{-\infty}^{y}(y-\eta)^{k}\exp\!\left\{e^{-g}-e^{-(x+\eta)}\right\}\,\mathrm{d}\eta
=0eλλkk!eλgexp[eg(eλ1)]dλ\displaystyle=\int_{0}^{\infty}e^{-\lambda}\frac{\lambda^{k}}{k!}e^{\lambda-g}\exp\!\left[-e^{-g}(e^{\lambda}-1)\right]\,\mathrm{d}\lambda
=0eλλkk!(Λgdλ),\displaystyle=\int_{0}^{\infty}e^{-\lambda}\frac{\lambda^{k}}{k!}\operatorname{\mathbb{P}{}}(\Lambda_{g}\in\mathrm{d}\lambda), (6.6)

where we have recalled (6.1) at (6.6). This completes the proof of (i) and thus the proof of the corollary. ∎

Acknowledgements.

We thank Ao Sun for providing a simplified proof of Lemma 3.2(i), and Daniel Q. Naiman and Ao Sun for valuable assistance in producing the three figures. We thank Svante Janson for helpful discussions about the details of this paper. We are also grateful for discussions with Persi Diaconis, Hsien-Kuei Hwang, Daniel Q. Naiman, Robin Pemantle, Ao Sun, and Nicholas Wormald. Last but not least, we thank two anonymous reviewers for many helpful suggestions.

References

  • [1] Patrick Billingsley. Probability and measure. Wiley Series in Probability and Statistics. John Wiley & Sons, Inc., Hoboken, NJ, 2012. Anniversary edition [of MR1324786], With a foreword by Steve Lalley and a brief biography of Billingsley by Steve Koppes.
  • [2] Graham Brightwell. Random kk-dimensional orders: width and number of linear extensions. Order, 9(4):333–342, 1992.
  • [3] James Allen Fill. Breaking bivariate records. Combin. Probab. Comput., 30(1):105–123, 2021.
  • [4] James Allen Fill and Daniel Q. Naiman. Generating Pareto records, 2019. arXiv:1901.05621.
  • [5] James Allen Fill and Daniel Q. Naiman. The Pareto record frontier. Electron. J. Probab., 25:1–24, 2020.
  • [6] Günter Last and Mathew Penrose. Lectures on the Poisson process, volume 7 of Institute of Mathematical Statistics Textbooks. Cambridge University Press, Cambridge, 2018.
  • [7] Peter Winkler. Random orders. Order, 1(4):317–331, 1985.