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

Filling random cycles

Fedor Manin Department of Mathematics, UCSB, Santa Barbara, California, USA [email protected]
Abstract.

We compute the asymptotic behavior of the average-case filling volume for certain models of random Lipschitz cycles in the unit cube and sphere. For example, we estimate the minimal area of a Seifert surface for a model of random knots first studied by Millett. This is a generalization of the classical Ajtai–Komlós–Tusnády optimal matching theorem from combinatorial probability. The author hopes for applications to the topology of random links, random maps between spheres, and other models of random geometric objects.

1. Introduction

1.1. Main results

This paper introduces a new kind of average-case isoperimetric inequality. Given a kk-cycle ZZ on ([0,1]n,[0,1]n)([0,1]^{n},\partial[0,1]^{n}), in any of a number of geometric measure theory senses, its filling volume FV(Z)FV(Z) is the minimal mass of a chain whose boundary is ZZ. The well-known Federer–Fleming isoperimetric inequality [FF] states that for all kk-cycles ZZ,

FV(Z)Cn,kmass(Z)k+1kandFV(Z)Cn,kmass(Z).FV(Z)\leq C_{n,k}\operatorname{mass}(Z)^{\frac{k+1}{k}}\quad\text{and}\quad FV(Z)\leq C_{n,k}\operatorname{mass}(Z).

However, one might expect that most cycles of given mass are much easier to fill.

Unfortunately, as explained to the author by Robert Young, a geometrically meaningful probability measure on the space of all cycles of mass N\leq N may be too much to ask for. The issue is one of picking a scale: say we are trying to build a random 1-Lipschitz curve in a finite-dimensional space. If the curve is to fluctuate randomly at scale ε\varepsilon, then over time 11 it will only travel a distance on the order of ε\sqrt{\varepsilon}. Thus there is no way of ensuring random behavior in a scale-free way. This idea of decomposing a finite-mass cycle into pieces at different scales can be made precise using the notion of a corona decomposition, as in [Jones] (in dimension 1) and [Young] (in higher dimensions).

On the other hand, there are a number of ways of, and a number of motivations for, building random “space-filling” cycles of mass O(N)O(N) which look essentially trivial on balls of radius N1/nN^{-1/n}. Our main theorems characterize three models of this form which exhibit similar isoperimetric behavior, and we hypothesize that this behavior should be generic for models which are random at many scales, including the largest—an idea which may have a precise Fourier-analytic formulation.

This isoperimetric behavior is described in codimension dd by the function

AKTd(N)={Nif d=1NlogNif d=2N(d1)/dif d3.\operatorname{AKT}_{d}(N)=\begin{cases}\sqrt{N}&\text{if }d=1\\ \sqrt{N\log N}&\text{if }d=2\\ N^{(d-1)/d}&\text{if }d\geq 3.\\ \end{cases}
Theorem A.

Let ZZ be a kk-cycle on SnS^{n} obtained by sampling NN oriented great kk-spheres independently from the uniform distribution on the oriented Grassmannian Gr~k+1(n+1)\widetilde{\operatorname{Gr}}_{k+1}(\mathbb{R}^{n+1}). Then there are constants C>c>0C>c>0 depending on nn and kk such that

(1.1) cAKTnk(N)<𝔼(FV(Z))<CAKTnk(N).c\operatorname{AKT}_{n-k}(N)<\mathbb{E}(FV(Z))<C\operatorname{AKT}_{n-k}(N).

Moreover, FV(Z)FV(Z) is concentrated around its mean: there are constants C1,C2>0C_{1},C_{2}>0 depending on nn and kk such that

(1.2) for every r>0[|FV(Z)𝔼(FV(Z))|r]C1exp(C2Nr)\text{for every $r>0$, }\mathbb{P}[\lvert FV(Z)-\mathbb{E}(FV(Z))\rvert\geq r]\leq C_{1}\exp(-C_{2}\sqrt{N}r)

From (1.2) we see that the spread of the distribution around the mean is at most on the order of N\sqrt{N}; in codimensions d=nk2d=n-k\geq 2, this is small compared to the mean:

for every ε>0|FV(Z)𝔼(FV(Z))|𝔼(FV(Z))ε with high probability as N.\text{for every $\varepsilon>0$, }\frac{\lvert FV(Z)-\mathbb{E}(FV(Z))\rvert}{\mathbb{E}(FV(Z))}\leq\varepsilon\text{ with high probability as }N\to\infty.

If a model of random codimension-dd cycles of mass O(N)O(N) satisfies (1.1) and (1.2), we say it exhibits AKT statistics, in honor of Ajtai, Komlós, and Tusnády, who discovered this phenomenon in the case of zero-cycles.

By rescaling the picture, we can make this result more interpretable. Let R=N1/(nk)R=N^{1/(n-k)}. Then the corresponding process in the nn-sphere of radius RR generates a cycle of mass Θ(Rn)\Theta(R^{n}) which is evenly spread throughout the sphere, so that a 1-ball intersects one of the great kk-spheres in ZZ on average. With that rescaling, the mass of an optimal filling becomes

{Θ(RnR)if k=n1Θ(RnlogR)if k=n2Θ(Rn)if kn3.\left\{\begin{array}[]{l l}\Theta(R^{n}\sqrt{R})&\text{if }k=n-1\\ \Theta(R^{n}\sqrt{\log R})&\text{if }k=n-2\\ \Theta(R^{n})&\text{if }k\leq n-3.\\ \end{array}\right.

Informally speaking, to meet its match, the average point in ZZ has to travel a distance Θ(R)\Theta(\sqrt{R}) (in codimension 1), Θ(logR)\Theta(\sqrt{\log R}) (in codimension 2), or Θ(1)\Theta(1) (otherwise) times the distance to its closest neighbor.

We also prove similar results for the cube:

Theorem B.

Let ZZ be a relative kk-cycle on ([0,1]n,[0,1]n)([0,1]^{n},\partial[0,1]^{n}) obtained by sampling NN planes independently from the uniform distribution on the space YY of oriented kk-planes which intersect [0,1]n[0,1]^{n} nontrivially. Then ZZ exhibits AKT statistics.

There may be reasonable disagreement as to which distribution is the uniform one in this context; to prove (1.1) it suffices to require that it be uniform on each subset of YY (isometric to two copies of a polytope) consisting of parallel planes, but to prove (1.2) we also need to assume that it behaves reasonably with respect to the manifold structure on YY (for example, is a positive density or has finite support).

In fact, the only thing used here about ZZ is that almost all of its “slices” along coordinate kk-planes consist of O(N)O(N) independent uniformly distributed points. This means that there are a number of other possible models that can be fit into this framework. However, the following requires a separate proof:

Theorem C.

Let {MN}\{M_{N}\} be a sequence of kk-dimensional oriented pseudomanifolds with NN vertices and at most LL simplices incident to any given simplex. Let ZZ be a kk-cycle on [0,1]n[0,1]^{n} obtained by sending each vertex of MNM_{N} to a uniformly random point in [0,1]n[0,1]^{n} and extending linearly. Then ZZ exhibits AKT statistics.

In the context of this theorem, the constants in (1.1) depend on nn, kk, and LL, but not on the shapes of the pseudomanifolds (which can therefore also be randomized). The case k=1k=1, n=3n=3 describes the “random jump” model of random knots and links introduced by Millett [Mil]. Moreover, by a theorem of Hardt and Simon [HS], the optimal filling of such a knot or link (after a slight rounding of corners) is a C1C^{1} embedded surface. In particular:

Corollary 1.3.

For some C>c>0C>c>0, the minimal Seifert surface of a knot produced using NN random jumps has area between cNlogNc\sqrt{N\log N} and CNlogNC\sqrt{N\log N} with high probability.

1.2. Motivation

The methods we use to prove Theorems A, B, and C can be easily extended to other i.i.d. samples of simple shapes on various spaces. However, the investigation is mainly motivated by the desire to analyze topological invariants of random geometric objects such as links and maps. Models of such objects tend to produce random cycles which are similarly trivial at small scales, but are more difficult to sample because they cannot be easily written in terms of i.i.d. parameters.

Random knots and links

There have been a number of proposed models of random knots and links; see [E-Z] for a detailed survey. Several of these models are “spatial” in the sense that they produce random knotted curves in space, and one supposes that these may exhibit AKT statistics for filling area. As mentioned above, we show this for Millett’s random jump model, but it may also be true for random polygonal walks with shorter segments as well as random grid walks, perhaps with some restrictions on segment length.

Given two random curves in a certain model, one may want to understand the distribution of their linking number. Since this will usually be zero on average, the first interesting question is about the second moment. Linking number can be computed as the intersection number of one curve with a filling of the other, thus one may expect that two random curves of length NN which exhibit AKT statistics have expected squared linking number NlogN\sim N\log N or NlogN\sim N\sqrt{\log N}.

However, this is not the case for the Millett model: the second moment of the linking number between two random jump curves of length NN is N\sim N [ABD+, FK]. Similarly, one may take the setup of Theorem A for k=1k=1 and n=3n=3 as a model of a random link and try to understand the total linking number, that is, the sum of the signed linking numbers of all pairs of circles. This is then the intersection number of the chain with its own filling. Here it is easy to see (as pointed out by Matthew Kahle) that the second moment of the distribution is once again N\sim N.

In both cases, this seeming incongruity perhaps boils down once again to the issue of multiple scales: random jump curves and great circles only “see” the largest scales, but the lower bound on filling volume in codimension 2 comes from looking on many different scales at once. One may perhaps get a different answer most easily by analyzing the linking number of an asymmetric model: a random jump curve and a random walk of total length NN made of smaller segments.

In [Tanaka, Marko], the second moment is computed for the linking number of two random walks; normalizing so that these walks have length NN and expected diameter 11, this second moment again becomes N\sim N. In this model, however, randomness happens at scale 1/N\sim 1/N, so it is not expected to exhibit AKT statistics.

Random maps

Another way of producing a random (framed) link is as the preimage of a generic point under a random map f:S3S2f:S^{3}\to S^{2}. In fact, the self-linking number of this link is the Hopf invariant of the map, which is itself a natural subject for investigation since it is a complete topological invariant of such maps.

One natural model of LL-Lipschitz random maps is a uniformly random simplicial map from a triangulation of S3S^{3} at scale L\sim L to a tetrahedron. The maximal self-linking number of such a map is Θ(L4)\Theta(L^{4}), cf. [Gro]; on the other hand, the heuristics above would suggest that the second moment of the linking number of the random model is between L3L^{3} and L3logLL^{3}\log L.

These ideas may have applications in topological and geometric data analysis, see [FKW].

1.3. Methods

The k=0k=0 cases of Theorems A and B are, up to minor adjustments, a classical theorem in combinatorial probability:

Theorem 1.4 (Ajtai, Komlós, and Tusnády [AKT]).

Let {X1,,XN}\{X_{1},\ldots,X_{N}\} and {Y1,,YN}\{Y_{1},\ldots,Y_{N}\} be two sets of independent, uniformly distributed random points in [0,1]d[0,1]^{d}, and let LL be the transportation cost between {Xi}\{X_{i}\} and {Yi}\{Y_{i}\}, that is, the total length of an optimal matching. Then there are constants 0<cd<Cd0<c_{d}<C_{d} such that with high probability,

cdAKTd(N)<L<CdAKTd(N).c_{d}\operatorname{AKT}_{d}(N)<L<C_{d}\operatorname{AKT}_{d}(N).

Since the original geometric proof in [AKT] of the most subtle case d=2d=2, this and related results have been proved many other times, often by applying Fourier analysis; see [BL2] for further references and [Tal3] for a detailed treatment of certain analytic approaches. Another beautiful geometric proof of the upper bound on the sphere is due to Holden, Peres, and Zhai [HPZ].

The proofs of Theorems A and B in general are obtained by applying the k=0k=0 results to (nk)(n-k)-dimensional slices of the cube and sphere. This is the reason that the results depend only on the codimension, and for the critical nature of codimension 2. The lower bound in (1.1) is obtained directly by integrating the lower bounds on these slices. The upper bound is obtained via a dual result on differential forms; this kind of technique was already used in [AKT] for the proof of the lower bound for the square. Finally, (1.2) is proved using the notion of concentration of measure due originally to Gromov and Milman [GM]; see [Led] for an extensive modern treatment.

Theorem C is proved similarly, except that slices no longer consist of i.i.d. points. Even this small amount of dependence complicates the argument considerably. We use ad hoc combinatorial arguments to overcome this, but one might hope to generalize, for example by applying a variant of Stein’s method, to a version of Theorem 1.4 in the presence of dependence (one approach, which only gives upper bounds, is discussed in [BL2, §5]).

Structure of the paper

Section 2 introduces necessary ideas and results from geometric measure theory, and Section 3 discusses the classical AKT theorem. In Sections 4 and 5, the upper and lower bounds in Theorems A and B are proved using tools that may generalize to other models of random cycles. In Section 6, we discuss the extra ideas needed to prove Theorem C. Finally, Section 7 discusses the concentration of the distributions in these theorems around their mean.

Acknowledgements

I would like to thank Matthew Kahle and Robert Young for a large number of helpful discussions over a span of three years. Yevgeny Liokumovich provided a crucial reference; Shmuel Weinberger asked a question which inspired Theorem C and gave other helpful comments. Finally, I would like to thank both Robert Young and the anonymous referee for calling out a great deal of sloppy and lazy writing in the initial draft. I was partially supported by NSF individual grant DMS-2001042.

2. Definitions and preliminaries

2.1. Cycles and currents

There are a number of useful ways to define chains and cycles from the point of view of topology and geometric measure theory. Algebraic topology typically uses singular kk-chains: formal linear combinations of continuous maps from the kk-simplex to a topological space XX (“singular simplices”). We will usually restrict our attention to Lipschitz simplices (that is, requiring the maps to be Lipschitz) on a Riemannian manifold MM. By Rademacher’s theorem, a Lipschitz simplex σ:ΔkM\sigma:\Delta^{k}\to M is differentiable almost everywhere and so has a well-defined volume or mass,

mass(σ)=ΔkσdvolM.\operatorname{mass}(\sigma)=\int_{\Delta^{k}}\sigma^{*}d\operatorname{vol}_{M}.

We can then extend by linearity to define the mass of a Lipschitz chain.

A more general notion of chain is that of a normal current. A kk-dimensional current on a manifold MM is simply a functional on (smooth) differential forms, which we think of as integration over the current. For example:

  • Every Lipschitz chain TT defines a current via ωTω\omega\mapsto\int_{T}\omega.

  • Every compactly supported (nk)(n-k)-form αΩnk(M)\alpha\in\Omega^{n-k}(M) defines a current via ωMαω\omega\mapsto\int_{M}\alpha\wedge\omega.

We will write the value of TT on ω\omega either as T(ω)T(\omega) or as Tω\int_{T}\omega, since currents should be thought of as generalized domains of integration. The boundary operator is defined via Stokes’ theorem: for a current TT,

T(ω)=T(dω).\partial T(\omega)=T(d\omega).

The mass of a kk-current TT on MM, which agrees with the same notion on Lipschitz chains, is defined to be

mass(T)=inf{T(ω):ωΩk(M) and ω=1}.\operatorname{mass}(T)=\inf\{T(\omega):\omega\in\Omega^{k}(M)\text{ and }\lVert\omega\rVert_{\infty}=1\}.

Here ω\lVert\omega\rVert_{\infty} is the supremum of the value of ω\omega over all frames of unit vectors. For a general current, the mass of course need not be finite. A current TT is normal if TT and T\partial T both have finite mass; in particular any cycle (current with empty boundary) of finite mass is normal.

2.2. Fillings and duality

Now, if SS is a normal current such that S=T\partial S=T, we call it a filling of TT. The filling volume of TT is

FV(T)=inf{mass(S)S=T},FV(T)=\inf\{\operatorname{mass}(S)\mid\partial S=T\},

which is always finite by the work of Federer and Fleming. The following is an instance of the Hahn–Banach theorem:

Proposition 2.1.

Let MM be a manifold. Then a normal kk-current TT in MM with T=0\partial T=0 has a filling of mass cc if and only if for every ωΩk(M)\omega\in\Omega^{k}(M) with dω1\lVert d\omega\rVert_{\infty}\leq 1, Tωc\int_{T}\omega\leq c.

More generally, for any closed set AMA\subset M, write Ωk(M,A)\Omega^{k}(M,A) for the vector space of forms whose restriction to AA is zero. Let TT be a normal kk-current with T\partial T supported on AA, that is, such that Tα=0\int_{\partial T}\alpha=0 for any (k1)(k-1)-form αΩk1(M,A)\alpha\in\Omega^{k-1}(M,A). Then TT has a filling relative to AA (that is, a (k+1)(k+1)-current SS such that ST\partial S-T is supported on AA) of mass cc if and only if for every ωΩk(M,A)\omega\in\Omega^{k}(M,A) with dω1\lVert d\omega\rVert_{\infty}\leq 1, Tωc\int_{T}\omega\leq c.

In other words, the filling volume of a cycle TT can be redefined as

FV(T)=sup{TωωΩk(M) such that dω1}FV(T)=\sup\bigl{\{}\textstyle{\int_{T}\omega}\mid\omega\in\Omega^{k}(M)\text{ such that }\lVert d\omega\rVert_{\infty}\leq 1\bigr{\}}

in both the absolute and the relative case. Our proofs of the upper bounds in Theorems A and B will be based on this proposition rather than constructing fillings directly.

Of course, knowing that a nice Lipschitz cycle has a filling which is a normal current is not very satisfying—after all, normal currents can still be very strange. Luckily, given a normal current filling, we can upgrade it to a Lipschitz chain (at the cost of multiplying the mass by a constant) using the following classical theorem:

Theorem 2.2 (Federer–Fleming deformation theorem [FF, Thm. 5.5]).

There is a constant ρ(k,n)=2n2k+2\rho(k,n)=2n^{2k+2} such that the following holds. Let TT be a normal current in 𝐍k(n)\mathbf{N}_{k}(\mathbb{R}^{n}). Then for every ε>0\varepsilon>0 we can write T=P+Q+ST=P+Q+\partial S, where

  1. (1)

    mass(P)ρ(k,n)mass(T)\operatorname{mass}(P)\leq\rho(k,n)\operatorname{mass}(T).

  2. (2)

    mass(Q)ερ(k,n)mass(T)\operatorname{mass}(Q)\leq\varepsilon\rho(k,n)\operatorname{mass}(\partial T).

  3. (3)

    mass(S)ερ(k,n)mass(T)\operatorname{mass}(S)\leq\varepsilon\rho(k,n)\operatorname{mass}(T).

  4. (4)

    PP is a polyhedral cycle which can be expressed as an \mathbb{R}-linear combination of kk-cells in the cubical unit lattice in n\mathbb{R}^{n}.

  5. (5)

    If TT is a Lipschitz chain, then so are QQ and SS.

  6. (6)

    If T\partial T is a Lipschitz chain, then so is QQ.

If TT is a normal current filling a Lipschitz chain T\partial T, then P+QP+Q is a Lipschitz chain filling TT whose mass is only greater by a multiplicative constant ρ(k,n)\rho(k,n).

It is not hard to upgrade the deformation theorem to manifolds, although the resulting constants will depend on the manifold and its metric; see for example [EPC+, Theorem 10.3.3].

2.3. Slicing

An important property of normal currents, introduced in [FF, §3], is the ability to take “slices” by hyperplanes to produce currents in lower dimensions. We follow the exposition of F. Morgan [Mor, 4.11], who follows Federer [Fed, §4.2.1].

Let u:Mu:M\to\mathbb{R} be a Lipschitz function on a manifold MM. Given a kk-current TT on MM and a differential rr-form ω\omega, define the (kr)(k-r)-current TωT\mathbin{\llcorner}\omega by

Tω(η)=T(ωη).T\mathbin{\llcorner}\omega(\eta)=T(\omega\wedge\eta).

In particular, this makes sense when ω\omega is a measurable function, for example the characteristic function χA\chi_{A} of a set AA. In that case we can write TA=TχAT\mathbin{\llcorner}A=T\mathbin{\llcorner}\chi_{A} for the restriction of TT to AA.

Given a Lipschitz function u:Mu:M\to\mathbb{R}, the slice of TT at u(x)=ru(x)=r is defined by

(2.3) T{u(x)=r}=(T){u(x)>r}(T{u(x)>r}).T\cap\{u(x)=r\}=(\partial T)\mathbin{\llcorner}\{u(x)>r\}-\partial(T\mathbin{\llcorner}\{u(x)>r\}).

If TT is a normal current, then T{u(x)=r}T\cap\{u(x)=r\} is a normal current for almost all rr. In addition, we have the following standard properties:

  1. (1)

    If TT is defined by integration over a (rectifiable) set XMX\subset M, then T{u(x)=r}T\cap\{u(x)=r\} is defined by integration over M{u(x)=r}M\cap\{u(x)=r\}.

  2. (2)

    T{u(x)=r}=(T{u(x)=r})\partial T\cap\{u(x)=r\}=-\partial(T\cap\{u(x)=r\}).

  3. (3)

    mass(T)1Lipumass(T{u(x)=r})𝑑r.\operatorname{mass}(T)\geq\frac{1}{\operatorname{Lip}u}\int_{-\infty}^{\infty}\operatorname{mass}(T\cap\{u(x)=r\})dr.

A quick calculation from the definitions yields an additional property:

Proposition 2.4.

If ω\omega is a (k1)(k-1)-form, then

T(duω)=(T{u(x)=r})(ω)𝑑r.T(du\wedge\omega)=-\int_{-\infty}^{\infty}(T\cap\{u(x)=r\})(\omega)dr.
Proof.

Start with the equalities

(Tu)(ω)\displaystyle\partial(T\mathbin{\llcorner}u)(\omega) =(T{u(x)>r})(ω)dr\displaystyle=\int_{-\infty}^{\infty}\partial(T\mathbin{\llcorner}\{u(x)>r\})(\omega)dr
(Tu)(ω)\displaystyle(\partial T\mathbin{\llcorner}u)(\omega) =((T){u(x)>r})(ω)𝑑r.\displaystyle=\int_{-\infty}^{\infty}((\partial T)\mathbin{\llcorner}\{u(x)>r\})(\omega)dr.

Subtract one from the other; then applying Stokes’ theorem and the Leibniz rule on the left and (2.3) on the right, we get the desired identity. ∎

In particular, by inductively slicing in different directions, we get the following:

Proposition 2.5.

Let 1kmn1\leq k\leq m\leq n, and let TT be an mm-dimensional current on [0,1]n[0,1]^{n}. Given x=(x1,,xk)k\vec{x}=(x_{1},\ldots,x_{k})\in\mathbb{R}^{k}, let PxP_{\vec{x}} be the plane {x}×nk\{\vec{x}\}\times\mathbb{R}^{n-k}. Then there are (mk)(m-k)-dimensional currents TPxT\cap P_{\vec{x}}, normal for almost all x\vec{x}, such that

(TPx)=TPx\displaystyle\partial(T\cap P_{\vec{x}})=\partial T\cap P_{\vec{x}}
mass(T)[0,1]kmass(TPx)𝑑x.\displaystyle\operatorname{mass}(T)\geq\int_{[0,1]^{k}}\operatorname{mass}(T\cap P_{\vec{x}})d\vec{x}.

In addition, given an mm-index I{1,,n}I\subset\{1,\ldots,n\} and a function f:[0,1]nf:[0,1]^{n}\to\mathbb{R},

Tf𝑑xI=(1)m[0,1]Ic(TPxf)𝑑x.\int_{T}fdx_{I}=(-1)^{m}\int_{[0,1]^{I^{c}}}\Bigl{(}\int_{T\cap P_{\vec{x}}}f\Bigr{)}d\vec{x}.

3. A variation on the Ajtai–Komlós–Tusnády theorem

The results of this paper are a generalization of Theorem 1.4. Properly, the theorem of Ajtai, Komlós, and Tusnády [AKT] is in the case n=2n=2; their paper also asserts the case n3n\geq 3, which is easy and later proved and extended in several directions by Talagrand [Tal1, Tal2]. The n=1n=1 case is elementary, and the proof along with a vast array of strengthenings and generalizations can be found in [BL1] by Bobkov and Ledoux. Here we need a slight variation.

Theorem 3.1.

Generate a cycle ZZ of mass NN in C0([0,1]n,[0,1]n)C_{0}([0,1]^{n},\partial[0,1]^{n}) by selecting NN independent, uniformly distributed points in [0,1]n×{+1,1}[0,1]^{n}\times\{+1,-1\}. Then there are constants 0<cn<Cn0<c_{n}<C_{n} such that

cnAKTn(N)𝔼(FV(Z))CnAKTn(N).c_{n}\operatorname{AKT}_{n}(N)\leq\mathbb{E}(FV(Z))\leq C_{n}\operatorname{AKT}_{n}(N).
Remark 3.2.

Suppose that DD is a Riemannian ball diffeomorphic to [0,1]n[0,1]^{n} and has a volume form. Then by the main theorem of [BMPR] (extending results of Moser [Moser] and Banyaga [Bany]), there is a diffeomorphism between the two which multiplies the volume form by a constant. Therefore Theorem 3.1 also holds with respect to Lebesgue measure on DD, with constants 0<cD<CD0<c_{D}<C_{D} depending on the ratio of the volumes and the bilipschitz constant of this diffeomorphism.

Moreover, given a smooth family of Riemannian balls, [BMPR] indicates that there is a smooth family of such diffeomorphisms. Therefore, if the family is compact, one can find uniform constants for the whole family.

Proof.

There are two differences here from the results as they are typically presented in the probability literature, where the problem consists of matching two sets of random points of the same cardinality: first, the number of positive and negative points may not match; second, we are allowed to match points to the boundary as well as to points of the opposite orientation.222In fact, a somewhat similar, but more complicated modification was studied by Shor [Shor]. We briefly explain how to modify the original proofs to deal with this.

Clearly, the possibility of matching to the boundary cannot make the upper bounds worse. Let’s say without loss of generality there are more positive points. To obtain the upper bound for n2n\geq 2, we may simply ignore some arbitrary set of “extra” positive points, matching all the others first. By the central limit theorem, the expected number of extra points is O(N)O(\sqrt{N}), so the extra mass generated by matching them all to the boundary of the cube does not change the asymptotic answer.

For the lower bound in the case n=2n=2, we use the same stratagem of ignoring the “extra” points to create a new cycle ZZ^{\prime} with an equal number of positive and negative points. From the original proof in [AKT], we know that there is a 1-Lipschitz function f:[0,1]2f:[0,1]^{2}\to\mathbb{R} which is zero on [0,1]2\partial[0,1]^{2} and such that ZfcNlogN\int_{Z^{\prime}}f\geq c\sqrt{N\log N} with high probability. Since with high probability the number of extra points is <<NlogN<<\sqrt{N\log N}, and the values of ff lie between 1/2-1/2 and 1/21/2, we also know that ZfcNlogN\int_{Z}f\geq c\sqrt{N\log N} with high probability.

The lower bound in the case n3n\geq 3 is easy to see: conditional on any distribution of the positive points, most negative points will be at distance cN1/n\geq cN^{-1/n} from every positive point and the boundary, where c>0c>0 is a constant depending on nn.

In the case n=1n=1, the filling is unique up to a constant: the unique filling FF supported away from zero has density 0xZ\int_{0}^{x}Z at x[0,1]x\in[0,1]. We use arguments found in [BL1, §3] to give estimates on 𝔼(massF)\mathbb{E}(\operatorname{mass}F).

The upper bound is a simple calculation:

𝔼(massF)\displaystyle\mathbb{E}(\operatorname{mass}F) =01𝔼(|0xZ|)𝑑x\displaystyle=\int_{0}^{1}\mathbb{E}\bigl{(}\bigl{\lvert}\textstyle{\int_{0}^{x}Z}\bigr{\rvert}\bigr{)}dx
01Var(0xZ)𝑑x=N2.\displaystyle\leq\int_{0}^{1}\sqrt{\operatorname{Var}(\textstyle{\int_{0}^{x}Z})}dx=\frac{\sqrt{N}}{2}.

The lower bound comes from the following classical fact, found in [BL1] as Lemma 3.4:

Lemma 3.3.

Given independent mean zero random variables ξ1,,ξN\xi_{1},\ldots,\xi_{N},

𝔼(|k=1Nξk|)122𝔼((k=1Nξk2)1/2).\mathbb{E}\biggl{(}\biggl{\lvert}\sum_{k=1}^{N}\xi_{k}\biggr{\rvert}\biggr{)}\geq\frac{1}{2\sqrt{2}}\mathbb{E}\biggl{(}\biggl{(}\sum_{k=1}^{N}\xi_{k}^{2}\biggr{)}^{1/2}\biggr{)}.

Let (Xk,σk)[0,1]×{+1,1}(X_{k},\sigma_{k})\in[0,1]\times\{+1,-1\} be the kkth chosen point. Then applying the lemma to ξk=σkχ{Xkx}\xi_{k}=\sigma_{k}\chi_{\{X_{k}\leq x\}}, we get

𝔼(|0xZ|)\displaystyle\mathbb{E}\bigl{(}\bigl{\lvert}\textstyle{\int_{0}^{x}Z}\bigr{\rvert}\bigr{)} 122𝔼((k=1Nξk2)1/2)\displaystyle\geq\frac{1}{2\sqrt{2}}\mathbb{E}\biggl{(}\biggl{(}\sum_{k=1}^{N}\xi_{k}^{2}\biggr{)}^{1/2}\biggr{)}
122(k=1N(𝔼(|ξk|))2)1/2=122Nx,\displaystyle\geq\frac{1}{2\sqrt{2}}\biggl{(}\sum_{k=1}^{N}(\mathbb{E}(|\xi_{k}|))^{2}\biggr{)}^{1/2}=\frac{1}{2\sqrt{2}}\sqrt{N}x,

and therefore 𝔼(massF)N/32\mathbb{E}(\operatorname{mass}F)\geq\sqrt{N/32}. ∎

4. Proof of the upper bound

To prove the upper bound in Theorems A and B, we will use Stokes’ theorem; that is, we use the fact that for a cycle ZCk(M,A)Z\in C_{k}(M,A),

(4.1) FV(Z)=sup{Zα:αΩk(M,A) such that dα=1}.FV(Z)=\sup\left\{\textstyle{\int_{Z}}\alpha:\alpha\in\Omega^{k}(M,A)\text{ such that }\lVert d\alpha\rVert_{\infty}=1\right\}.

In fact, since ZZ is a cycle, Zα\int_{Z}\alpha only depends on ω=dα\omega=d\alpha. To bound this quantity, we first note that any ωΩk+1([0,1]n)\omega\in\Omega^{k+1}([0,1]^{n}) can be decomposed into a sum of “basic” forms of the form

ωI(x)dxi1dxik+1,\omega_{I}(x)dx_{i_{1}}\wedge\cdots\wedge dx_{i_{k+1}},

where ωI\omega_{I} is a function n\mathbb{R}^{n}\to\mathbb{R}, for each subset {i1,,ik+1}{1,,n}\{i_{1},\ldots,i_{k+1}\}\subset\{1,\ldots,n\}.

Lemma 4.2.

For any exact form ωΩk+1([0,1]n,[0,1]n)\omega\in\Omega^{k+1}([0,1]^{n},\partial[0,1]^{n}) (resp., ωΩk+1([0,1]n)\omega\in\Omega^{k+1}([0,1]^{n})), there is a form αΩk([0,1]n,[0,1]n)\alpha\in\Omega^{k}([0,1]^{n},\partial[0,1]^{n}) (resp., αΩk([0,1]n)\alpha\in\Omega^{k}([0,1]^{n})) given by

α=I[n]|I|=kαI(x)dxI,\alpha=\sum_{\begin{subarray}{c}I\subset[n]\\ |I|=k\end{subarray}}\alpha_{I}(x)dx_{I},

such that dα=ωd\alpha=\omega, and for each II, αILip=dαICn,kω\lVert\alpha_{I}\rVert_{\operatorname{Lip}}=\lVert d\alpha_{I}\rVert_{\infty}\leq C_{n,k}\lVert\omega\rVert_{\infty}.

Proof.

We prove this by induction on nn and kk, keeping nkn-k constant. In the base case k=0k=0, we can take the function α\alpha to be the fiberwise integral 01ω\int_{0}^{1}\omega along one of the coordinates.

To do the inductive step in the relative case, we follow the usual proof of the Poincaré lemma with compact support, following [BottTu, §1.4]. Fix a smooth bump function ε:[0,1][0,1]\varepsilon:[0,1]\to[0,1] which is 0 near 0 and 1 near 1. By applying the lemma one dimension lower, we get a form ηΩk1([0,1]n1,[0,1]n1)\eta\in\Omega^{k-1}([0,1]^{n-1},\partial[0,1]^{n-1}) with dη=01ωd\eta=\int_{0}^{1}\omega and ηILipCn1,k1ω\lVert\eta_{I}\rVert_{\operatorname{Lip}}\leq C_{n-1,k-1}\lVert\omega\rVert_{\infty}. Then ω=dα\omega=d\alpha for

α=0tωε(xn)π(01ω)dε(xn)πη,\alpha={\textstyle\int_{0}^{t}\omega}-\varepsilon(x_{n})\pi^{*}({\textstyle\int_{0}^{1}\omega})-d\varepsilon(x_{n})\wedge\pi^{*}\eta,

where π\pi is the projection to the (n1)(n-1)-cube along xnx_{n}. Notice that

αI={dεdxηI{n}if nI0tωI{n}ε(x1)π(01ωI{n})otherwise.\alpha_{I}=\begin{cases}-{\displaystyle\frac{d\varepsilon}{dx}}\eta_{I\setminus\{n\}}&\text{if }n\in I\\ \int_{0}^{t}\omega_{I\cup\{n\}}-\varepsilon(x_{1})\pi^{*}(\int_{0}^{1}\omega_{I\cup\{n\}})&\text{otherwise.}\end{cases}

This gives us a bound on each αILip\lVert\alpha_{I}\rVert_{\operatorname{Lip}} in terms of the ωILip\lVert\omega_{I}\rVert_{\operatorname{Lip}} and ηILip\lVert\eta_{I}\rVert_{\operatorname{Lip}} as well as the derivatives of ε\varepsilon.

For the non-relative version, we follow the same proof, mutatis mutandis, taking

α=0tωπη.\alpha={\textstyle\int_{0}^{t}\omega}-\pi^{*}\eta.\qed

Here and in the next section, by a random kk-cycle we mean a random variable taking values in kk-cycles, that is, a measure on the space of kk-cycles. We follow the convention, common in probability theory, of blurring the distinction between a measure on a set of objects and an object randomly drawn from that measure.

Theorem 4.3.

Let ZZ be a random kk-cycle in ([0,1]n,[0,1]n)([0,1]^{n},\partial[0,1]^{n}) or in [0,1]n[0,1]^{n} which satisfies the condition that for some M>0M>0,

(4.4) 𝔼(FV(ZP))M\mathbb{E}(FV(Z\cap P))\leq M

for almost all (nk)(n-k)-planes PP parallel to one of the coordinate (nk)(n-k)-planes. Then

(4.5) 𝔼(FV(Z))(nk)Cn,kM,\mathbb{E}(FV(Z))\leq{n\choose k}C_{n,k}M,

where Cn,kC_{n,k} is the constant from Lemma 4.2.

Proof.

For a kk-form α\alpha, Zα\int_{Z}\alpha depends only on dαd\alpha. Therefore to estimate (4.1) it is enough to show that for any (k+1)(k+1)-form ω\omega with ω=1\lVert\omega\rVert_{\infty}=1, there is a kk-form α\alpha such that dα=ωd\alpha=\omega and ZαCn,kM\int_{Z}\alpha\leq C_{n,k}M.

By Lemma 4.2, we can choose

α=I[n]|I|=kαI(x)dxI\alpha=\sum_{\begin{subarray}{c}I\subset[n]\\ |I|=k\end{subarray}}\alpha_{I}(x)dx_{I}

such that dαICn,k\lVert d\alpha_{I}\rVert_{\infty}\leq C_{n,k} for every II. Then for α\alpha ranging over all these choices of antidifferentials,

FV(Z)\displaystyle FV(Z) =supαZα=supαI[n]|I|=k[0,1]IcZPuαI𝑑u\displaystyle=\sup_{\alpha}\int_{Z}\alpha=\sup_{\alpha}\sum_{\begin{subarray}{c}I\subset[n]\\ |I|=k\end{subarray}}\int_{[0,1]^{I^{c}}}\int_{Z\cap P_{u}}\alpha_{I}du
I[n]|I|=k[0,1]Ic(supαZPuαI)𝑑uI[n]|I|=k[0,1]IcCn,kFV(ZPu).\displaystyle\leq\sum_{\begin{subarray}{c}I\subset[n]\\ |I|=k\end{subarray}}\int_{[0,1]^{I^{c}}}\left(\sup_{\alpha}\int_{Z\cap P_{u}}\alpha_{I}\right)du\leq\sum_{\begin{subarray}{c}I\subset[n]\\ |I|=k\end{subarray}}\int_{[0,1]^{I^{c}}}C_{n,k}FV(Z\cap P_{u}).

By linearity of expectation,

𝔼(FV(Z))I[n]|I|=k[0,1]IcCn,k𝔼(FV(ZPu))(nk)Cn,kM.\mathbb{E}(FV(Z))\leq\sum_{\begin{subarray}{c}I\subset[n]\\ |I|=k\end{subarray}}\int_{[0,1]^{I^{c}}}C_{n,k}\mathbb{E}(FV(Z\cap P_{u}))\leq{n\choose k}C_{n,k}M.\qed
Proof of Theorem B, upper bound..

Let ZZ be a cycle in Ck([0,1]n,[0,1]n)C_{k}([0,1]^{n},\partial[0,1]^{n}) obtained by sampling NN i.i.d. planes from a distribution on the set of oriented kk-planes which intersect nontrivially with [0,1]n[0,1]^{n}, such that the distribution is uniform (with respect to Lebesgue measure on the corresponding polytope in nk\mathbb{R}^{n-k}) on each set of parallel planes.

This condition clearly implies that for every coordinate (nk)(n-k)-plane PP, ZPZ\cap P consists of at most NN i.i.d. positive and negative points with probability 1. Then Theorem 3.1 implies that (4.4) holds for ZZ with M=CnkAKTnk(N)M=C_{n-k}\operatorname{AKT}_{n-k}(N).

𝔼(FV(Z))2(nk)Cn,kCnAKTnk(N).\mathbb{E}(FV(Z))\leq 2{n\choose k}C_{n,k}C_{n}\operatorname{AKT}_{n-k}(N).\qed
Proof of Theorem A, upper bound..

We use the fact that the transverse intersection of an oriented great kk-sphere with an oriented great (nk)(n-k)-sphere is a pair of antipodal points with opposite orientations. Therefore, if ZZ is a cycle obtained by sampling NN oriented great kk-spheres independently from the uniform distribution, then for any great (nk)(n-k)-sphere PP, with probability 1

ZP=i=1N[xi][xi]Z\cap P=\sum_{i=1}^{N}[x_{i}]-[-x_{i}]

where the xix_{i} are i.i.d. uniform points on SnkS^{n-k}.

Consider SnS^{n} as a subset of n+1\mathbb{R}^{n+1}, with standard unit basis vectors e0,,ene_{0},\ldots,e_{n}. Let Ki±K_{i}^{\pm} be the preimage of the cube [R,R]n[-R,R]^{n} under central projection (that is, projection along lines through the origin) to the plane xi=±1x_{i}=\pm 1. If RR is large enough, the interiors of the Ki±K_{i}^{\pm} cover SnS^{n}. Each Ki+K_{i}^{+} is disjoint from its antipodal set KiK_{i}^{-}; therefore for any great (nk)(n-k)-sphere PP, with probability 1 ZPKi±Z\cap P\cap K_{i}^{\pm} consists of i.i.d. uniform points. By Remark 3.2,

𝔼(FV(ZPKi±))Cn,kAKTnk(N),\mathbb{E}(FV(Z\cap P\cap K_{i}^{\pm}))\leq C_{n,k}\operatorname{AKT}_{n-k}(N),

where ZPKi±Z\cap P\cap K_{i}^{\pm} is considered as a cycle in C0(PKi±,(PKi±))C_{0}(P\cap K_{i}^{\pm},\partial(P\cap K_{i}^{\pm})).

Note that central projection sends great spheres to hyperplanes. Therefore, by Theorem 4.3, for each ii and sign,

(4.6) 𝔼(FV(ZKi±))Cn,kAKTnk(N).\mathbb{E}(FV(Z\cap K_{i}^{\pm}))\leq C_{n,k}\operatorname{AKT}_{n-k}(N).

Set a partition of unity {φi±}\{\varphi_{i}^{\pm}\} subordinate to {Ki±}\{K_{i}^{\pm}\} which is invariant with respect to the involution, that is, such that φi+(x)=φi(x)\varphi_{i}^{+}(x)=\varphi_{i}^{-}(-x). To prove the theorem, it is enough, given a kk-form ωCk+1(Sn)\omega\in C_{k+1}(S^{n}) with ω=1\lVert\omega\rVert_{\infty}=1, to show that for some (and therefore every) αCk(Sn)\alpha\in C_{k}(S^{n}) with dα=ωd\alpha=\omega,

ZαCn,kAKTnk(N).\int_{Z}\alpha\leq C_{n,k}\operatorname{AKT}_{n-k}(N).

But note that

Zα=i=0n(ZKi+φi+α+ZKiφiα).\int_{Z}\alpha=\sum_{i=0}^{n}\Bigl{(}\int_{Z\cap K_{i}^{+}}\varphi_{i}^{+}\alpha+\int_{Z\cap K_{i}^{-}}\varphi_{i}^{-}\alpha\Bigr{)}.

Therefore it suffices to find an antidifferential and a bound separately for each φi±ω\varphi_{i}^{\pm}\omega. Therefore (4.6) suffices to prove the theorem. ∎

5. Proof of the lower bound

Theorem 5.1.

Let ZZ be a random Lipschitz kk-cycle in ([0,1]n,[0,1]n)([0,1]^{n},\partial[0,1]^{n}) such that for almost every kk-plane

Px={(x1,,xk)}×[0,1]nk[0,1]n,P_{\vec{x}}=\{(x_{1},\ldots,x_{k})\}\times[0,1]^{n-k}\subset[0,1]^{n},

the slice ZPxZ\cap P_{\vec{x}} satisfies

𝔼(FV(ZPx))p(x)\mathbb{E}(FV(Z\cap P_{\vec{x}}))\geq p(\vec{x})

where p:[0,1]k[0,)p:[0,1]^{k}\to[0,\infty) is an L1L^{1} function. Then

𝔼(FV(Z))[0,1]kp(x)𝑑x.\mathbb{E}(FV(Z))\geq\int_{[0,1]^{k}}p(\vec{x})d\vec{x}.
Proof.

Let UU be a normal current filling ZZ such that mass(U)FV(Z)+ε\operatorname{mass}(U)\leq FV(Z)+\varepsilon, for any ε>0\varepsilon>0. Then for almost all PxP_{\vec{x}}, there is a slice UPxU\cap P_{\vec{x}} which fills ZPxZ\cap P_{\vec{x}}, and

mass(UPx)FV(ZPx).\operatorname{mass}(U\cap P_{\vec{x}})\geq FV(Z\cap P_{\vec{x}}).

By Proposition 2.5,

FV(Z)+εmass(U)[0,1]kmass(UPx)𝑑x[0,1]kFV(ZPx)𝑑x.FV(Z)+\varepsilon\geq\operatorname{mass}(U)\geq\int_{[0,1]^{k}}\operatorname{mass}(U\cap P_{\vec{x}})d\vec{x}\geq\int_{[0,1]^{k}}FV(Z\cap P_{\vec{x}})d\vec{x}.

Since this is true for every ε>0\varepsilon>0, and by linearity of expectation,

𝔼(FV(Z))[0,1]k𝔼(FV(ZPx))𝑑x[0,1]kp(x)𝑑x.\mathbb{E}(FV(Z))\geq\int_{[0,1]^{k}}\mathbb{E}(FV(Z\cap P_{\vec{x}}))d\vec{x}\geq\int_{[0,1]^{k}}p(\vec{x})d\vec{x}.\qed
Proof of Theorem B, lower bound..

Let ZZ be a cycle in Ck([0,1]n,[0,1]n)C_{k}([0,1]^{n},\partial[0,1]^{n}) obtained by sampling NN i.i.d. planes from a distribution on the set of oriented kk-planes which intersect nontrivially with [0,1]n[0,1]^{n}, such that the distribution is uniform (with respect to Lebesgue measure on the corresponding polytope in nk\mathbb{R}^{n-k}) on each set of parallel planes. Assume, perhaps by permuting coordinates, that this distribution is not concentrated on planes of the form

Px=(x1,,xk)×nk.P_{\vec{x}}=(x_{1},\ldots,x_{k})\times\mathbb{R}^{n-k}.

As in the proof of the upper bound, it follows that for every PxP_{\vec{x}}, ZPxZ\cap P_{\vec{x}} consists of i.i.d. positive and negative points with probability 1. Moreover, the probability of a random plane PP intersecting PxP_{\vec{x}} inside [0,1]n[0,1]^{n} depends only on the direction of PP and not on x\vec{x}. Thus

𝔼(mass(ZPx))cN,\mathbb{E}(\operatorname{mass}(Z\cap P_{\vec{x}}))\geq cN,

where cc depends on the distribution but not on x\vec{x}.Thus by Theorems 3.1 and 5.1,

𝔼(FV(Z))12CnkAKTnk(cN).\mathbb{E}(FV(Z))\geq\frac{1}{2}C_{n-k}\operatorname{AKT}_{n-k}(cN).\qed
Proof of Theorem A, lower bound..

Let ZZ be a cycle in Ck(Sn)C_{k}(S^{n}) obtained by sampling NN independent uniformly distributed great kk-spheres. It suffices to show that for some compact submanifold KSnK\subset S^{n},

FV(ZK)Cn,kAKTnk(N)FV(Z\cap K)\geq C_{n,k}\operatorname{AKT}_{n-k}(N)

where ZKZ\cap K is considered as a cycle in Ck(K,K)C_{k}(K,\partial K).

Recall that for any great (nk)(n-k)-sphere PP, with probability 1

ZP=i=1N[xi][xi]Z\cap P=\sum_{i=1}^{N}[x_{i}]-[-x_{i}]

where the xix_{i} are i.i.d. uniform points on SnkS^{n-k}. Let TSnT\subset S^{n} be a great kk-sphere and TT^{\prime} the (nk)(n-k)-sphere consisting of points farthest from TT. We use Nε(U)N_{\varepsilon}(U) to indicate the ε\varepsilon-neighborhood of the set UU; then SnNπ/4(T)=Nπ/4(T)¯S^{n}\setminus N_{\pi/4}(T^{\prime})=\overline{N_{\pi/4}(T)} deformation retracts to TT along the orthogonal retraction ρ:Nπ/4(T)¯T\rho:\overline{N_{\pi/4}(T)}\to T. We let K=ρ1(K)K=\rho^{-1}(K^{\prime}) where KK^{\prime} is some closed ball in TT which does not include any point and its antipode.

Notice that KK is foliated by equal-volume patches of great (nk)(n-k)-spheres PuP_{u} which retract to points uKu\in K^{\prime}, and also does not include any point and its antipode. By Remark 3.2, there is a bilipschitz diffeomorphism from KK to [0,1]n[0,1]^{n} which sends each PuP_{u} to a plane of the form

(x1,,xk)×nk(x_{1},\ldots,x_{k})\times\mathbb{R}^{n-k}

in a volume-preserving way (up to a constant). Therefore, for each uKu\in K^{\prime},

𝔼(FV(ZPuK))cAKTnk(cN),\mathbb{E}(FV(Z\cap P_{u}\cap K))\geq c\operatorname{AKT}_{n-k}(cN),

and applying Theorem 5.1, we obtain the result. ∎

6. Proof of Theorem C

Before we prove Theorem C, we must give a more precise statement.

By an oriented kk-pseudomanifold we mean a kk-dimensional simplicial complex MM with the following properties:

  • It is pure, i.e. every simplex is contained in an kk-dimensional simplex.

  • Every kk-simplex comes with an orientation such that the sum of all the oriented kk-simplices is a cycle in Ck(M)C_{k}(M).

Note that this is considerably wider than the usual definition: it is just enough so that if MM is equipped with the standard simplexwise metric, any Lipschitz map from MM to a metric space XX defines a Lipschitz kk-cycle in XX.

We say MM has geometry bounded by LL if every kk-simplex intersects at most LL others.

With these definitions, we restate Theorem C:

Theorem.

Let MM be an oriented kk-pseudomanifold with NN vertices and geometry bounded by LL. Let ZZ be a kk-cycle on [0,1]n[0,1]^{n} obtained by sending each vertex of MM to a uniformly random point in [0,1]n[0,1]^{n} and extending linearly. Then there are constants C>c>0C>c>0 depending on nn and kk such that

(6.1) cL1AKTnk(N)<𝔼(FV(Z))<CLAKTnk(N).cL^{-1}\operatorname{AKT}_{n-k}(N)<\mathbb{E}(FV(Z))<CL\operatorname{AKT}_{n-k}(N).

The concentration result will be proved in the next section.

As with Theorem B, the proof is a direct application of Theorems 4.3 and 5.1. To apply these theorems, we need to understand the filling volumes of slices of ZZ, which is more complicated in this case because while the points are identically distributed, they are not entirely independent. We establish the upper bound in Lemma 6.3; this depends only on the fact that every point is independent of all but a constant number of others. For the lower bound in Theorem 6.13, the argument is more subtle: even if every point is correlated with only one other, such pairs could be very close and have opposite signs; then the least filling would be much smaller than for independent points. Accordingly, we have to show that most correlated points are still far apart. Together, these two bounds complete the proof.

In this section as before, fix the notation

Px={x}×[0,1]nk[0,1]n,x[0,1]k.P_{\vec{x}}=\{\vec{x}\}\times[0,1]^{n-k}\subset[0,1]^{n},\qquad\vec{x}\in[0,1]^{k}.

We start by analyzing the slice ZPxZ\cap P_{\vec{x}}.

Lemma 6.2.

Let x[0,1]k\vec{x}\in[0,1]^{k}. Then the slice ZPxZ\cap P_{\vec{x}} is the sum of NN random 0-chains ζ1,,ζN\zeta_{1},\ldots,\zeta_{N} which are identically distributed on {±[y]:y[0,1]nk}{0}\{\pm[y]:y\in[0,1]^{n-k}\}\cup\{0\} according to a distribution μx\mu_{\vec{x}} depending on kk and x\vec{x}. Moreover:

  1. (i)

    μx\mu_{\vec{x}} is invariant with respect to the involution sending a chain ζ\zeta to ζ-\zeta;

  2. (ii)

    μxC(n,k)μLebesgue\mu_{\vec{x}}\leq C(n,k)\mu_{\text{Lebesgue}} on [0,1]nk[0,1]^{n-k}.

  3. (iii)

    Each ζi\zeta_{i} is independent of all but at most LL other ζj\zeta_{j}.

Since the distribution of ZZ is invariant under permuting coordinates, this holds for any (nk)(n-k)-dimensional slice in a coordinate direction.

Proof.

The distribution in question is the intersection of a random linear kk-simplex in [0,1]n[0,1]^{n} with PxP_{\vec{x}}. Property (i) is obvious from this, and (iii) follows since a pair of ζi\zeta_{i} are independent whenever the two corresponding simplices do not intersect. To see (ii), consider the function

Fx:(([0,1]n)k+1,μLebesgue)({±[y]:y[0,1]nk}{0},μx)F_{\vec{x}}:\bigl{(}([0,1]^{n})^{k+1},\mu_{\text{Lebesgue}}\bigr{)}\to\bigl{(}\{\pm[y]:y\in[0,1]^{n-k}\}\cup\{0\},\mu_{\vec{x}}\bigr{)}

sending each linear kk-simplex to its intersection with PxP_{\vec{x}}. This function is measure-preserving by definition, and its restriction to

K=Fx1{[y]:y[0,1]nk}K=F_{\vec{x}}^{-1}\{[y]:y\in[0,1]^{n-k}\}

is 1-Lipschitz. Therefore, by the coarea formula, the density function of μx\mu_{\vec{x}} is given by the [n(k+1)(nk)][n(k+1)-(n-k)]-dimensional Hausdorff measure of point preimages. Thus it is enough to bound H(n+1)k(Fx1(y))H_{(n+1)k}(F_{\vec{x}}^{-1}(\vec{y})) for each y\vec{y}.

Let TT be the set of linear kk-simplices with vertices in [1,1]n[-1,1]^{n} which pass through 0\vec{0}, and let

Tx=(T+(x,0))([0,1]k×[1,1]nk)k+1.T_{\vec{x}}=(T+(\vec{x},\vec{0}))\cap([0,1]^{k}\times[-1,1]^{n-k})^{k+1}.

(Here each vertex is translated by (x,0)(\vec{x},\vec{0}).) Notice that

Fx1(y)Tx+(0,y).F_{\vec{x}}^{-1}(\vec{y})\subset T_{\vec{x}}+(\vec{0},\vec{y}).

All these translates are disjoint and their union is a subset of ([0,1]k×[1,2]nk)k([0,1]^{k}\times[-1,2]^{n-k})^{k}. Therefore, again by the coarea formula,

H(n+1)k(Tx)3(k+1)(nk).H_{(n+1)k}(T_{\vec{x}})\leq 3^{(k+1)(n-k)}.

This completes the proof of (ii). ∎

Condition (iii) gives a dependency graph of degree L\leq L between the ζi\zeta_{i}. This graph has an (L+1)(L+1)-coloring, giving a partition of {0,,N}\{0,\ldots,N\} into L+1L+1 disjoint subsets I1,,InI_{1},\ldots,I_{n} such that for iIji\in I_{j}, the ζi\zeta_{i} are i.i.d.

6.1. Upper bound

This lemma gives the upper bound to plug into Theorem 4.3:

Lemma 6.3.

For every x[0,1]k\vec{x}\in[0,1]^{k},

𝔼(FV(ZPx))(L+1)Cn,kAKTnk(N).\mathbb{E}(FV(Z\cap P_{\vec{x}}))\leq(L+1)C_{n,k}\operatorname{AKT}_{n-k}(N).
Proof.

We estimate 𝔼(FV(ZPx))\mathbb{E}(FV(Z\cap P_{\vec{x}})) by separately considering each summand

Z(x,j)=iIjζi,j=1,,L+1.Z(\vec{x},j)=\sum_{i\in I_{j}}\zeta_{i},\qquad j=1,\ldots,L+1.

These summands consist of i.i.d. negative and positive points.

Write νx\nu_{\vec{x}} for the probability measure on [0,1]nk[0,1]^{n-k} given by

νx(A)=μx{[y]:yA}μx{[y]:y[0,1]nk}.\nu_{\vec{x}}(A)=\frac{\mu_{\vec{x}}\{[\vec{y}]:\vec{y}\in A\}}{\mu_{\vec{x}}\{[\vec{y}]:\vec{y}\in[0,1]^{n-k}\}}.

If ζ\zeta is a random 0-cycle in [0,1]nk[0,1]^{n-k} with NN positive and NN negative points distributed according to νx\nu_{\vec{x}}, then the AKT upper bound holds for ζ\zeta: by [BL2, equation (12)], for some constant CnkC_{n-k} independent of the measure,

(6.4) 𝔼(FV(ζ))CnkAKTnk(N).\mathbb{E}(FV(\zeta))\leq C_{n-k}\operatorname{AKT}_{n-k}(N).

To reduce to this situation, we note that while ZPxZ\cap P_{\vec{x}} is a cycle (and therefore has an equal number of negative and positive points), Z(x,j)Z(\vec{x},j) may not be. We produce cycles Z~(x,j)\tilde{Z}(\vec{x},j) for j=1,,L+1j=1,\ldots,L+1 by adding up to NN additional i.i.d. points distributed according to νx\nu_{\vec{x}}. We add each point to Z~(x,j)\tilde{Z}(\vec{x},j) for two different jj, with opposite signs, so that

j=1L+1Z~(x,j)=j=1L+1Z(x,j)=ZPx.\sum_{j=1}^{L+1}\tilde{Z}(\vec{x},j)=\sum_{j=1}^{L+1}Z(\vec{x},j)=Z\cap P_{\vec{x}}.

Each Z~(x,j)\tilde{Z}(\vec{x},j) is a 0-cycle consisting of at most NN positive and NN negative i.i.d. points. Therefore, by (6.4),

𝔼(FV(ZPx))(L+1)Cn,kAKTnk(N).\mathbb{E}(FV(Z\cap P_{\vec{x}}))\leq(L+1)C_{n,k}\operatorname{AKT}_{n-k}(N).\qed

6.2. Lower bound

For the lower bound, we begin by showing that correlated points in ZPxZ\cap P_{\vec{x}} are usually not very close; this relationship is summarized in Lemma 6.5. We use this as an ingredient in the proof of Theorem 6.13, which retraces some of the steps in the original proof of the AKT theorem.

In many places, we refer to the function FxF_{\vec{x}} and set TxT_{\vec{x}} defined in the proof of Lemma 6.2. Also, given a subset A[0,1]nkA\subseteq[0,1]^{n-k}, we write

[A]={[y]:yA}and±[A]={±[y]:yA}[A]=\{[y]:y\in A\}\qquad\text{and}\qquad{\pm[A]}=\{\pm[y]:y\in A\}

for the respective sets of 0-dimensional chains.

Lemma 6.5.

Assume that x[1/4,3/4]k\vec{x}\in[1/4,3/4]^{k}. Let 0<<1/20<\ell<1/2 and let ζ\zeta and ζ\zeta^{\prime} be random variables whose values are the intersections with PxP_{\vec{x}} of two kk-simplices Δ\Delta and Δ\Delta^{\prime} of MM whose vertices, some of which are shared, are chosen uniformly at random from [0,1]n[0,1]^{n}. Let Q[1/4,3/4]nkQ\subset[1/4,3/4]^{n-k} be a cube of side length \ell. Then

(6.6) [ζ±[Q]ζ[Q]]Cn,k.\mathbb{P}\bigl{[}\zeta^{\prime}\in\pm[Q]\mid\zeta\in[Q]\bigr{]}\leq C_{n,k}\sqrt{\ell}.
Remark 6.7.

A more careful analysis based on the same principle shows that

[ζ±[Q]ζ[Q]]{Cn,k|log|if nk=1,Cn,kotherwise.\mathbb{P}\bigl{[}\zeta^{\prime}\in\pm[Q]\mid\zeta\in[Q]\bigr{]}\leq\begin{cases}C_{n,k}\ell\lvert\log\ell\rvert&\text{if }n-k=1,\\ C_{n,k}\ell&\text{otherwise}.\end{cases}
Proof.

The idea is this: suppose that Δ\Delta and Δ\Delta^{\prime} share a (k1)(k-1)-face Δ0\Delta_{0}, and let ww and ww^{\prime} be the non-shared vertices of Δ\Delta and Δ\Delta^{\prime}. If the intersections of Δ\Delta and Δ\Delta^{\prime} with PxP_{\vec{x}} are close to each other, then either the angle between Δ\Delta and Δ\Delta^{\prime} is small (forcing ww^{\prime} to be close to the kk-plane containing Δ\Delta), or Δ0\Delta_{0} is close to PxP_{\vec{x}}. We would like to show that neither of these happens too often.

We will do this case in detail; the analysis when Δ\Delta and Δ\Delta^{\prime} share a lower-dimensional face is similar.

First, we show that Δ0\Delta_{0} is not very often too close to PxP_{\vec{x}}. This is easy to establish globally; the tricky bit is showing that it’s true after conditioning on ζ\zeta being supported in QQ. Let ρx(Δ0)\rho_{\vec{x}}(\Delta_{0}) be the distance from Δ0\Delta_{0} to the hyperplane {x}×nk\{\vec{x}\}\times\mathbb{R}^{n-k}. The following lemma will be proved later.

Lemma 6.8.

Let x[1/4,3/4]k\vec{x}\in[1/4,3/4]^{k} and let A[1/4,3/4]nkA\subseteq[1/4,3/4]^{n-k} be a set of positive measure. Let Δ\Delta be a simplex with vertices chosen uniformly at random from [0,1]n[0,1]^{n}, and let Δ0\Delta_{0} be its 0th face. Then

[ρx(Δ0)rFx(Δ)[A]]Cn,kr.\mathbb{P}\bigl{[}\rho_{\vec{x}}(\Delta_{0})\leq r\mid F_{\vec{x}}(\Delta)\in[A]\bigr{]}\leq C_{n,k}r.

In particular,

(6.9) [ρx(Δ0)ζ[Q]]Cn,k.\mathbb{P}\bigl{[}\rho_{\vec{x}}(\Delta_{0})\leq\sqrt{\ell}\mid\zeta\in[Q]\bigr{]}\leq C_{n,k}\sqrt{\ell}.

Now we must show that when ρx(Δ0)>\rho_{\vec{x}}(\Delta_{0})>\sqrt{\ell}, ζ\zeta^{\prime} doesn’t very often land near ζ\zeta. The point is that the difference depends on the angle between Δ\Delta and Δ\Delta^{\prime}, and the distribution of this angle is not too concentrated anywhere; this is true for any given Δ0\Delta_{0}. Fix Δ0\Delta_{0} with ρ(Δ0)\rho(\Delta_{0})\geq\sqrt{\ell}, and let U(Δ0,Q)[0,1]nU(\Delta_{0},Q)\subseteq[0,1]^{n} be the set of points ww^{\prime} such that ζ±[Q]\zeta^{\prime}\in\pm[Q]. Note that U(Δ0,{z})U(\Delta_{0},\{z\}) is contained in the intersection of a kk-plane with [0,1]n(k+1)[0,1]^{n(k+1)} and hence its kk-dimensional Hausdorff norm is at most some Cn,kC_{n,k}. Now we would like to use the coarea formula to integrate with respect to zQz\in Q. For this we need the following estimate, to be proved later:

Lemma 6.10.

Given a linear kk-simplex Δ([0,1]n)k+1\Delta\in([0,1]^{n})^{k+1} such that at least one of its (k1)(k-1)-faces is at distance at least rr from PxP_{\vec{x}},

det([(DFx)Δ]T(DFx)Δ)Cn,krnk.\sqrt{\det\bigl{(}[(DF_{\vec{x}})_{\Delta}]^{T}(DF_{\vec{x}})_{\Delta}\bigr{)}}\geq C_{n,k}r^{n-k}.

Then by the coarea formula, for fixed Δ0\Delta_{0},

(k)nk2vol(U(Δ0,Q))Cn,kvol(Q)\left(\frac{\ell}{k}\right)^{\frac{n-k}{2}}\operatorname{vol}(U(\Delta_{0},Q))\leq C_{n,k}\operatorname{vol}(Q)

and therefore

vol(U(Δ0,Q))Cn,knk2.\operatorname{vol}(U(\Delta_{0},Q))\leq C_{n,k}\ell^{\frac{n-k}{2}}.

Integrating this over the domain in [0,1]kn[0,1]^{kn} containing all values of Δ0\Delta_{0} such that ρx(Δ0)\rho_{\vec{x}}(\Delta_{0})\geq\sqrt{\ell}, we see that

(6.11) [ζ±[Q]ζ[Q],ρ(Δ0)]Cn,knk2.\mathbb{P}\bigl{[}\zeta^{\prime}\in\pm[Q]\mid\zeta\in[Q],\rho(\Delta_{0})\geq\sqrt{\ell}\bigr{]}\leq C_{n,k}\ell^{\frac{n-k}{2}}.

Together, (6.9) and (6.11) imply (6.6). ∎

Now we prove the lemmas.

PxP_{\vec{x}}Δyi\Delta y_{i}DkD\leq\sqrt{k}drd\geq r
Figure 1. When a vertex moves by Δyi\Delta y_{i}, ΔPx\Delta\cap P_{\vec{x}} moves by (d/D)Δyi(d/D)\Delta y_{i}.
Proof of Lemma 6.10.

From Figure 1, we see that for every 1ink1\leq i\leq n-k, there is a unit vector v\vec{v} such that (DFx)Δ(v)=cei(DF_{\vec{x}})_{\Delta}(\vec{v})=c\vec{e}_{i}, for c>r/kc>r/\sqrt{k}. Therefore

det([(DFx)Δ]T(DFx)Δ)(rk)nk.\sqrt{\det\bigl{(}[(DF_{\vec{x}})_{\Delta}]^{T}(DF_{\vec{x}})_{\Delta}\bigr{)}}\geq\left(\frac{r}{\sqrt{k}}\right)^{n-k}.\qed
Proof of Lemma 6.8.

Let XX be the event that ρx(Δ)r\rho_{\vec{x}}(\Delta)\leq r, and let YY be the event that ΔPx[Q]\Delta\cap P_{\vec{x}}\in[Q]. Bayes’ rule states that

(XY)=(YX)(X)/(Y).\mathbb{P}(X\mid Y)=\mathbb{P}(Y\mid X)\mathbb{P}(X)/\mathbb{P}(Y).

We will bound (XY)\mathbb{P}(X\mid Y) by giving upper bounds for (YX)\mathbb{P}(Y\mid X) and (X)\mathbb{P}(X) and a lower bound for (Y)\mathbb{P}(Y).

We first show that

(6.12) (YX)3(nk)(k+1)vol(A).\mathbb{P}(Y\mid X)\leq 3^{(n-k)(k+1)}\operatorname{vol}(A).

In fact, in this inequality, XX may be any constraint on the first kk coordinates of each vertex of Δ\Delta (note that ρx(Δ)\rho_{\vec{x}}(\Delta) depends only on those coordinates).

We start by fixing the first kk coordinates of each vertex. Given a simplex Δ\Delta with vertices in n\mathbb{R}^{n}, let π(Δ)\pi(\Delta) be the projection onto the first kk coordinates, and fix a simplex Δπ\Delta_{\pi} with vertices in [0,1]k[0,1]^{k}. Let U([1,1]nk)k+1U\subset([-1,1]^{n-k})^{k+1} contain the last nkn-k coordinates of simplices in π1(Δπ)\pi^{-1}(\Delta_{\pi}) whose intersection with PxP_{\vec{x}} is [(x,0)][(\vec{x},\vec{0})]. Write U+AU+A to mean the set of translates of simplices in UU by points in AA. Notice that

  1. (1)

    U+[0,1]nk([1,2]nk)k+1U+[0,1]^{n-k}\subseteq([-1,2]^{n-k})^{k+1},

  2. (2)

    the volume of U+AU+A is proportional to vol(A)\operatorname{vol}(A), and

  3. (3)

    U+{y}U+\{\vec{y}\} contains F1([y])π1(Δπ)F^{-1}([\vec{y}])\cap\pi^{-1}(\Delta_{\pi}).

Therefore, the probability that a given point of ([0,1]nk)k+1([0,1]^{n-k})^{k+1} is contained in U+AU+A is at most 3(nk)(k+1)vol(A)3^{(n-k)(k+1)}\operatorname{vol}(A), and this in turn bounds

vol(F1([A])π1(Δπ)).\operatorname{vol}(F^{-1}([A])\cap\pi^{-1}(\Delta_{\pi})).

Integrating over possible values of Δπ\Delta_{\pi} gives (6.12).

Now we show that (X)Cn,kr\mathbb{P}(X)\leq C_{n,k}r. Choosing kk points in k\mathbb{R}^{k} uniformly at random induces a probability measure on the set of (k1)(k-1)-planes, whose density at a plane PP is proportional to vol(P[0,1]k)k\operatorname{vol}(P\cap[0,1]^{k})^{k}. This density is bounded above by some Cn,kC_{n,k}, and so the set of planes whose distance from x\vec{x} is at most rr has measure Cn,kr\leq C_{n,k}r.

Finally we must give a lower bound for (Y)\mathbb{P}(Y), that is, the volume of

ΣY={Δ:Fx(Δ)[A]}[0,1]n(k+1).\Sigma_{Y}=\{\Delta:F_{\vec{x}}(\Delta)\in[A]\}\subseteq[0,1]^{n(k+1)}.

For this, we note that when x\vec{x} and y\vec{y} have all coordinates in [1/4,3/4][1/4,3/4],

Fx1([y])Tx([1/4,1/4]n+{(x,0)})+{(0,y)}.F_{\vec{x}}^{-1}([\vec{y}])\supseteq T_{\vec{x}}\cap([-1/4,1/4]^{n}+\{(\vec{x},\vec{0})\})+\{(\vec{0},\vec{y})\}.

This is a translate of a set which is independent of x\vec{x} and y\vec{y}. Taking all the translates with respect to vectors in AA, we see that ΣY\Sigma_{Y} contains a set whose volume depends linearly on vol(A)\operatorname{vol}(A) and is easily seen to be positive.

Thus overall we get (XY)Cn,kr\mathbb{P}(X\mid Y)\leq C_{n,k}r. ∎

Finally we have the tools we need to prove the lower bound for Theorem C.

Theorem 6.13.

For every x[1/4,3/4]k\vec{x}\in[1/4,3/4]^{k},

𝔼(FV(ZPx))cn,kL1AKTnk(N).\mathbb{E}(FV(Z\cap P_{\vec{x}}))\geq c_{n,k}L^{-1}\operatorname{AKT}_{n-k}(N).
Proof.

Write d=nkd=n-k. We are trying to find a lower bound for the expected filling length of a certain distribution on 0-cycles in [0,1]d[0,1]^{d} which is defined as a sum of a large number of identically distributed points. This is very similar to the original AKT theorem, in which the points are, in addition, independent and uniformly distributed. In [AKT], the lower bound is proved using the dual definition of filling volume given in (4.1), which we restate here for a 0-cycle Z0Z_{0} in [0,1]d[0,1]^{d}:

FV(Z0)=sup{Z0f:[0,1]d such that Lip(f)1}.FV(Z_{0})=\sup\bigl{\{}\textstyle{\int_{Z_{0}}}f:[0,1]^{d}\to\mathbb{R}\text{ such that }\operatorname{Lip}(f)\leq 1\bigr{\}}.

For the case d=2d=2, Ajtai, Komlós and Tusnády construct a NlogN\sqrt{N\log N}-Lipschitz function ff whose integral over Z0Z_{0} is usually at least cNlogNcN\log N. The filling volume FV(Z0)FV(Z_{0}) is then bounded below by the ratio of the integral to the Lipschitz constant.

When d=2d=2, we construct ff the same way as in [AKT], but on a modified point set in order to overcome two issues:

  1. (1)

    The points of ZPxZ\cap P_{\vec{x}} are not uniformly distributed. We use a bilipschitz rescaling of the domain to make sure that they are.

  2. (2)

    The points of ZPxZ\cap P_{\vec{x}} are not independent. We instead construct the function ff to have a large integral over a large independent subset, and then use Lemma 6.5 to show that the integral over the remaining points is small.

For d=1d=1 and d3d\geq 3, we use the same reduction steps, but the function ff is somewhat simpler. We start by more precisely describing the common argument and then give the detailed construction of ff for each case.

Let ζi\zeta_{i} be the chain-valued random variables corresponding to intersections of kk-simplices of ZZ with PxP_{\vec{x}}. Recall that ζi=±[vi]\zeta_{i}=\pm[v_{i}] or {0}\{0\}, with vi[0,1]dv_{i}\in[0,1]^{d} identically distributed according to a density which is bounded below on [1/4,3/4]d[1/4,3/4]^{d}. Moreover, this bound is uniform with respect to x[1/4,3/4]k\vec{x}\in[1/4,3/4]^{k}. Thus, following Remark 3.2, there is a uniformly bilipschitz family of diffeomorphisms

φx:[1/4,3/4]d[0,1]d\varphi_{\vec{x}}:[1/4,3/4]^{d}\to[0,1]^{d}

which send this density to a constant times the standard volume form. In particular, Lemma 6.5 still holds for the ζi\zeta_{i} after applying the diffeomorphism. We now write ζi\zeta_{i} for φx(ζi)\varphi_{\vec{x}}(\zeta_{i}).

In each case, consider an (L+1)(L+1)-coloring I1,,IL+1I_{1},\ldots,I_{L+1} of the dependency graph between the ζi\zeta_{i}, with the colored subsets ordered from largest to smallest. Note that each of the ζi\zeta_{i} is correlated with ζi\zeta_{i^{\prime}} for at most LL values of ii^{\prime}.

We write Zj=iIjζiZ_{j}=\sum_{i\in I_{j}}\zeta_{i}. In each case, we show that Z1Z_{1} is hard to fill by constructing a Lipschitz function ff such that

Z1fcn,kLAKTnk(N)Lipf with high probability.\int_{Z_{1}}f\geq\frac{c_{n,k}}{L}\operatorname{AKT}_{n-k}(N)\operatorname{Lip}f\text{ with high probability.}

Then we show that 𝔼(Zjf)\mathbb{E}(\int_{Z_{j}}f) for each j1j\neq 1 is small enough that it does not affect the overall asymptotics.

In each case, the function ff is constructed as a sum of simpler functions. Given a cube Q[0,1]dQ\subseteq[0,1]^{d}, let ΔQ:[0,1]d\Delta_{Q}:[0,1]^{d}\to\mathbb{R} be the function supported on QQ whose graph is a symmetric pyramid with base QQ and height 1. We also write Δvr\Delta^{r}_{v} for the cube in the lattice of side length rr which contains v[0,1]dv\in[0,1]^{d}. The function ff will consist of a sum of scaled copies of ΔQ\Delta_{Q}, each reflecting the “imbalance” of positive and negative points on the cube QQ. The main difference between different dimensions is the scale of these cubes: for d3d\geq 3, the cubes are at the smallest scale (comparable to the average distance between neighboring points), for d=1d=1 they are at the largest scale (comparable to 1), and for d=2d=2 we use cubes at many scales, as in the original proof of [AKT].

In each case, we construct ff by means of an auxiliary function

g(x)=iI1gζi(x)g(x)=\sum_{i\in I_{1}}g_{\zeta_{i}}(x)

(in the case d=2d=2, each ζi\zeta_{i} is associated to many summands gζirg^{r}_{\zeta_{i}} at different scales, which we consider separately). This gg may not satisfy the desired upper bound on the Lipschitz constant, so we remove some of the summands where they are too concentrated to produce ff. Whenever ζi\zeta_{i} is independent from ζi\zeta_{i^{\prime}}, 𝔼(ζigζi(x))=0\mathbb{E}\bigl{(}\int_{\zeta_{i^{\prime}}}g_{\zeta_{i}}(x)\bigr{)}=0, and so for every j1j\neq 1 we can write

|𝔼(Zjg)|iIj{|𝔼(ζigζi)|ζi is correlated with ζi}.\bigl{\lvert}\mathbb{E}\bigl{(}{\textstyle\int_{Z_{j}}}g\bigr{)}\bigr{\rvert}\leq\sum_{i\in I_{j}}\sum\{\bigl{\lvert}\mathbb{E}\bigl{(}{\textstyle\int_{\zeta_{i}}}g_{\zeta_{i^{\prime}}}\bigr{)}\bigr{\rvert}\mid\zeta_{i^{\prime}}\text{ is correlated with }\zeta_{i}\}.

By Lemma 6.5, this correlation is not too high, and therefore we can bound each of the summands. By the construction of ff, this also bounds |𝔼(Zjf)|\bigl{\lvert}\mathbb{E}\bigl{(}\int_{Z_{j}}f\bigr{)}\bigr{\rvert}.

If we tune everything correctly, we get that for each j1j\neq 1,

|𝔼(Zjf)|12L𝔼(Z1f),\bigl{\lvert}\mathbb{E}\bigl{(}{\textstyle\int_{Z_{j}}}f\bigr{)}\bigr{\rvert}\leq\frac{1}{2L}\mathbb{E}\bigl{(}{\textstyle\int_{Z_{1}}}f\bigr{)},

giving a lower bound on 𝔼(Zf)\mathbb{E}\bigl{(}\int_{Z}f\bigr{)}.

Case d3d\geq 3

We split the cube [0,1]d[0,1]^{d} into N\sim N subcubes of side length rN1/dr\approx N^{-1/d}, and let

g(x)\displaystyle g(x) =iI1gi(x)=iI1±rΔvir(x) where ζi=±[vi],\displaystyle=\sum_{i\in I_{1}}g_{i}(x)=\sum_{i\in I_{1}}\pm r\Delta^{r}_{v_{i}}(x)\text{ where }\zeta_{i}=\pm[v_{i}],
f(x)\displaystyle f(x) =t1,,td=0r11sign(Z1χQr(t1,,td))rΔQr(t1,,td)(x),\displaystyle=\sum_{t_{1},\ldots,t_{d}=0}^{r^{-1}-1}\operatorname{sign}\Bigl{(}\int_{Z_{1}}\chi_{Q_{r}(t_{1},\ldots,t_{d})}\Bigr{)}r\Delta_{Q_{r}(t_{1},\ldots,t_{d})}(x),

where Qr(t1,,td)Q_{r}(t_{1},\ldots,t_{d}) is the cube with side length rr whose vertex closest to the origin is (rt1,,rtd)(rt_{1},\ldots,rt_{d}). Note that ff is 22-Lipschitz.

We can think of the number of points landing in each subcube as NN independent λ=1\lambda=1 Poisson processes which we stop once their sum reaches roughly

(ζi[0,1]d0)|I1|.\mathbb{P}(\zeta_{i}\cap[0,1]^{d}\neq 0)\lvert I_{1}\rvert.

By the law of large numbers, the stopping time will be very close to

t=N1(ζi=±[y],y[0,1]d)|I1|t=N^{-1}\mathbb{P}(\zeta_{i}=\pm[y],y\in[0,1]^{d})\lvert I_{1}\rvert

and very nearly NtetNte^{-t} of the subcubes will contain exactly one point. Of these, with high probability, at least (1/2)3dNtet(1/2)\cdot 3^{-d}\cdot Nte^{-t} will be contained in the middle third of their subcube. Therefore,

Z1fcn,kLNd1d with high probability.\int_{Z_{1}}f\geq\frac{c_{n,k}}{L}\cdot N^{\frac{d-1}{d}}\text{ with high probability}.

On the other hand, given j1j\neq 1, iIji\in I_{j}, and iI1i^{\prime}\in I_{1} such that ζi\zeta_{i} is correlated with ζi\zeta_{i^{\prime}}, Lemma 6.5 tells us that

|𝔼(ζigi)|Cn,kr3/2\bigl{\lvert}\mathbb{E}\bigl{(}\textstyle{\int_{\zeta_{i}}}g_{i^{\prime}}\bigr{)}\bigr{\rvert}\leq C_{n,k}r^{3/2}

and therefore

|𝔼(Zjf)|iIjLCn,kr3/2LCn,kNd1d12d.\bigl{\lvert}\mathbb{E}\bigl{(}\textstyle{\int_{Z_{j}}}f\bigr{)}\bigr{\rvert}\leq\sum_{i\in I_{j}}LC_{n,k}r^{3/2}\leq LC_{n,k}N^{\frac{d-1}{d}-\frac{1}{2d}}.

Since this is small compared to Z1f\int_{Z_{1}}f, this shows that

𝔼(ZPxf)cn,kLNd1dLipffor large enough N.\mathbb{E}\Bigl{(}\int_{Z\cap P_{\vec{x}}}f\Bigr{)}\geq\frac{c_{n,k}}{L}N^{\frac{d-1}{d}}\operatorname{Lip}f\qquad\text{for large enough }N.

Case d=2d=2

This case is broadly similar, but we build the function ff in a more complicated way, following the original proof of [AKT]. For an integer rr, let QstrQ^{r}_{st} be the square of side length 2r2^{-r} whose lower left corner is at (s2r,t2r)(s\cdot 2^{-r},t\cdot 2^{-r}), and write Δstr=ΔQstr\Delta^{r}_{st}=\Delta_{Q^{r}_{st}}. We write

g(x,y)\displaystyle g(x,y) =r=10.1logNs,t=02r1gstr(x,y)=r=10.1logNs,t=02r1Δstr(x,y)Z1Δstr\displaystyle=\sum_{r=1}^{0.1\log N}\sum_{s,t=0}^{2^{r}-1}g^{r}_{st}(x,y)=\sum_{r=1}^{0.1\log N}\sum_{s,t=0}^{2^{r}-1}\Delta^{r}_{st}(x,y)\int_{Z_{1}}\Delta^{r}_{st}
=r=10.1logNiI1gζir(x,y)\displaystyle=\sum_{r=1}^{0.1\log N}\sum_{i\in I_{1}}g^{r}_{\zeta_{i}}(x,y)
=r=10.1logNiI1Δvir(x,y)ζiΔvirwhere ζi=±[vi].\displaystyle=\sum_{r=1}^{0.1\log N}\sum_{i\in I_{1}}\Delta^{r}_{v_{i}}(x,y)\int_{\zeta_{i}}\Delta^{r}_{v_{i}}\qquad\text{where }\zeta_{i}=\pm[v_{i}].

Notice that Z1gstr\int_{Z_{1}}g^{r}_{st} is always nonnegative: roughly speaking, it measures the square of the “imbalance” of positive and negative points in QstrQ^{r}_{st}. In particular, it’s not hard to see that 𝔼(Z1gstr)=cn,k22rN\mathbb{E}(\int_{Z_{1}}g^{r}_{st})=c_{n,k}\cdot 2^{-2r}N, and therefore 𝔼(Z1g)=cn,kNlogN\mathbb{E}(\int_{Z_{1}}g)=c_{n,k}N\log N.

On the other hand, the derivative of gg is O(NlogN)O(\sqrt{N\log N}) on average, but can be much larger in some places. To remedy this, Ajtai, Komlós, and Tusnády introduced a “stopping time” rule, building ff as the sum of some, but not all of the gstrg^{r}_{st}. We do not need to give the exact definition, remarking only that the function ff satisfies

(6.14) 𝔼(Z1f)\displaystyle\mathbb{E}\bigl{(}\textstyle{\int_{Z_{1}}}f\bigr{)} cn,kLNlogN\displaystyle\geq\frac{c_{n,k}}{L}N\log N
(6.15) Lipf\displaystyle\operatorname{Lip}f Cn,kNlogN.\displaystyle\leq C_{n,k}\sqrt{N\log N}.

Now, for j1j\neq 1,

|𝔼(Zjf)|iIjr=10.1logN{|𝔼(ζigζir)|ζi is correlated with ζi}.\bigl{\lvert}\mathbb{E}\bigl{(}{\textstyle\int_{Z_{j}}}f\bigr{)}\bigr{\rvert}\leq\sum_{i\in I_{j}}\sum_{r=1}^{0.1\log N}\sum\bigl{\{}\bigl{\lvert}\mathbb{E}\bigl{(}{\textstyle\int_{\zeta_{i}}}g^{r}_{\zeta_{i^{\prime}}}\bigr{)}\bigr{\rvert}\mid\zeta_{i^{\prime}}\text{ is correlated with }\zeta_{i}\bigr{\}}.

By Lemma 6.5, the value of each term of this triple sum is O(2r/2)O(2^{-r/2}). Therefore,

|𝔼(Zjf)|Cn,kLNr=10.1logN2r/2(1+2)Cn,kLN.\bigl{\lvert}\mathbb{E}\bigl{(}{\textstyle\int_{Z_{j}}}f\bigr{)}\bigr{\rvert}\leq C_{n,k}LN\sum_{r=1}^{0.1\log N}2^{-r/2}\leq(1+\sqrt{2})C_{n,k}LN.

Combining this with (6.14) and (6.15), we see that

𝔼(ZPxf)cn,kLNlogNLipffor large enough N.\mathbb{E}\Bigl{(}\int_{Z\cap P_{\vec{x}}}f\Bigr{)}\geq\frac{c_{n,k}}{L}\sqrt{N\log N}\operatorname{Lip}f\qquad\text{for large enough }N.

Case d=1d=1

We split the interval [0,1][0,1] into RR equal regions, with RR to be determined later. Write Δs=Δ[s/R,(s+1)/R]\Delta_{s}=\Delta_{[s/R,(s+1)/R]}, and let

g(x)=s=0R1Δs(x)Z1χ[sR,s+1R]=iI1gζi(x)=iI1±Δyi1/R(x)where ζi=±[yi].g(x)=\sum_{s=0}^{R-1}\Delta_{s}(x){\textstyle\int_{Z_{1}}}\chi_{\left[\frac{s}{R},\frac{s+1}{R}\right]}=\sum_{i\in I_{1}}g_{\zeta_{i}}(x)=\sum_{i\in I_{1}}\pm\Delta^{1/R}_{y_{i}}(x)\qquad\text{where }\zeta_{i}=\pm[y_{i}].

We obtain the desired function ff by replacing Z1χ[sR,s+1R]\int_{Z_{1}}\chi_{\left[\frac{s}{R},\frac{s+1}{R}\right]} with

hs=sign(Z1χ[sR,s+1R])min{|Z1χ[sR,s+1R]|,Cn,kNR}h_{s}=\operatorname{sign}\biggl{(}\int_{Z_{1}}\chi_{\left[\frac{s}{R},\frac{s+1}{R}\right]}\biggr{)}\min\biggl{\{}\biggl{\lvert}\int_{Z_{1}}\chi_{\left[\frac{s}{R},\frac{s+1}{R}\right]}\biggr{\rvert},C_{n,k}\sqrt{\frac{N}{R}}\biggr{\}}

for some sufficiently large Cn,kC_{n,k}. Then LipfCn,kNR\operatorname{Lip}f\leq C_{n,k}\sqrt{NR} and Z1fcn,kLN\int_{Z_{1}}f\geq\frac{c_{n,k}}{L}N.

On the other hand, for j1j\neq 1 and any iIji\in I_{j} and iI1i^{\prime}\in I_{1} such that ζi\zeta_{i} and ζi\zeta_{i^{\prime}} are correlated, by Lemma 6.5, |𝔼(ζigζi)|Cn,kR1/2\bigl{\lvert}\mathbb{E}\bigl{(}{\textstyle\int_{\zeta_{i}}}g_{\zeta_{i^{\prime}}}\bigr{)}\bigr{\rvert}\leq C_{n,k}R^{-1/2}, and therefore

|𝔼(Zjf)|Cn,kLNR1/2.\bigl{\lvert}\mathbb{E}\bigl{(}{\textstyle\int_{Z_{j}}}f\bigr{)}\bigr{\rvert}\leq C_{n,k}LNR^{-1/2}.

For some large enough RR, depending on nn and kk but not on NN,

|𝔼(Zjf)|12L2𝔼(Z1f).\bigl{\lvert}\mathbb{E}\bigl{(}{\textstyle\int_{Z_{j}}}f\bigr{)}\bigr{\rvert}\leq\frac{1}{2L^{2}}\mathbb{E}\bigl{(}{\textstyle\int_{Z_{1}}}f\bigr{)}.

Thus 𝔼(ZPxf)Cn,kNLipf\mathbb{E}\bigl{(}\int_{Z\cap P_{\vec{x}}}f\bigr{)}\geq C_{n,k}\sqrt{N}\operatorname{Lip}f, completing the proof. ∎

7. Concentration of measure

In this section, we show that when nk2n-k\geq 2, the size of the filling tends to concentrate around its mean. That is, we show that (1.2) holds in the case of Theorems A, B, and C. We first prove this in the case of Theorem 3.1. The main tool is the concentration of measure in high-dimensional balls, an idea due to Gromov and Milman [GM] and of wide importance in probability theory [Led]. We follow the exposition due to Bobkov and Ledoux [BL1, §7.1] which covers the 1-dimensional case; the higher-dimensional cases are essentially the same although they do not seem to appear explicitly in the literature.

Theorem 7.1.

Let ZZ be a random cycle in C0([0,1]n,[0,1]n)C_{0}([0,1]^{n},\partial[0,1]^{n}) as in Theorem 3.1. Then for every r>0r>0,

[|FV(Z)𝔼(FV(Z))|r]C1exp(C2r/N)\mathbb{P}[\lvert FV(Z)-\mathbb{E}(FV(Z))\rvert\geq r]\leq C_{1}\exp\bigl{(}-C_{2}r/\sqrt{N}\bigr{)}

for universal constants C1,C2>0C_{1},C_{2}>0.

In particular, the standard deviation of FV(Z)FV(Z) is at most O(N)O(\sqrt{N}). In other words, for n2n\geq 2, FV(Z)/𝔼(FV(Z))FV(Z)/\mathbb{E}(FV(Z)) converges to 1 as NN\to\infty.

Proof.

Equip X=C0([0,1]n,[0,1]n)X=C_{0}([0,1]^{n},\partial[0,1]^{n}) with the metric

dFV(Z,Z)=FV(ZZ)d_{FV}(Z,Z^{\prime})=FV(Z-Z^{\prime})

and let E=[1,1]×[0,1]n1E=[-1,1]\times[0,1]^{n-1}. Define ζ0:EX\zeta_{0}:E\to X by

ζ0(±x1,x2,,xn)=±[(x1,x2,,xn)]\zeta_{0}(\pm x_{1},x_{2},\ldots,x_{n})=\pm[(x_{1},x_{2},\ldots,x_{n})]

and ζ:(EN,dEucl)X\zeta:(E^{N},d_{\text{Eucl}})\to X by

ζ(v1,,vN)=i=1Nζ0(vi).\zeta(v_{1},\ldots,v_{N})=\sum_{i=1}^{N}\zeta_{0}(v_{i}).

This map is N\sqrt{N}-Lipschitz since when every point moves by a tiny amount ε\varepsilon, the distance is Nε\sqrt{N}\varepsilon in the domain and NεN\varepsilon in the range.

Define the concentration function of a metric measure space (M,d,μ)(M,d,\mu) of total measure 1 to be

concM(r)=sup{1μ(Nr(A))μ(A)1/2},r>0,\operatorname{conc}_{M}(r)=\sup\{1-\mu(N_{r}(A))\mid\mu(A)\geq 1/2\},\qquad r>0,

where Nr(A)N_{r}(A) is the rr-neighborhood of the set AA. The key observation of Gromov and Milman [GM, Thm. 4.1] is that

concM(r)34eln(3/2)λ1r,\operatorname{conc}_{M}(r)\leq\frac{3}{4}e^{-\ln(3/2)\lambda_{1}r},

where λ1\lambda_{1} is the first nonzero eigenvalue of the Laplacian on MM. Since the spectrum of a product of manifolds is the sum of its spectra, λ1\lambda_{1} is constant on powers of MM. The map ζ\zeta is measure-preserving, so it follows that

concX(r)34exp(ln(3/2)λ1r/N).\operatorname{conc}_{X}(r)\leq\frac{3}{4}\exp\bigl{(}-\ln(3/2)\lambda_{1}r/\sqrt{N}\bigr{)}.

Therefore, for any 11-Lipschitz function u:Xu:X\to\mathbb{R},

[|u(Z)median(u)|r]32exp(ln(3/2)λ1r/N).\mathbb{P}[|u(Z)-\operatorname{median}(u)|\geq r]\leq\frac{3}{2}\exp\bigl{(}-\ln(3/2)\lambda_{1}r/\sqrt{N}\bigr{)}.

By Chebyshev’s inequality, the same, modulo constants, holds for the mean (see also [Led, Prop. 1.10]). ∎

To adapt this proof for the case of Theorems A, B, and C, we just have to change the space EE: take

Esphere\displaystyle E_{\text{sphere}} =Gr~k(n)\displaystyle=\widetilde{Gr}_{k}(\mathbb{R}^{n}) for Theorem A
Ecube\displaystyle E_{\text{cube}} ={affine k-planes PnP[0,1]n}\displaystyle=\{\text{affine $k$-planes }P\subset\mathbb{R}^{n}\mid P\cap[0,1]^{n}\neq\emptyset\} for Theorem B
Eknot\displaystyle E_{\text{knot}} =[0,1]n\displaystyle=[0,1]^{n} for Theorem C.\displaystyle\text{for Theorem \ref{knot}}.

In the first two cases, the map ζ\zeta is constructed as before. In the last case, for a kk-pseudomanifold MM with vertex set M0M^{0}, ζM:Eknot|M0|Zk([0,1]n)\zeta_{M}:E_{\text{knot}}^{\lvert M^{0}\rvert}\to Z_{k}([0,1]^{n}) sends (v1,,v|M0|)(v_{1},\ldots,v_{|M^{0}|}) to the image of the linear immersion of MM with vertices (v1,,v|M0|)(v_{1},\ldots,v_{|M^{0}|}).

In each case, it is easy to see that if the space of kk-cycles is given the filling volume metric, then ζ\zeta is N\sqrt{N}-Lipschitz. Therefore, the rest of the proof is identical to that of Theorem 7.1.

References

  • [ABD+] J. Arsuaga, T. Blackstone, Y. Diao, E. Karadayi, and M. Saito, Linking of uniform random polygons in confined spaces, J. Phys. A 40 (2007), no. 9, 1925–1936.
  • [AKT] M. Ajtai, J. Komlós, and G. Tusnády, On optimal matchings, Combinatorica 4 (1984), no. 4, 259–264.
  • [Bany] A. Banyaga, Formes-volume sur les variétés à bord, Enseign. Math. (2) 20 (1974), 127–131.
  • [BL1] S. Bobkov and M. Ledoux, One-dimensional empirical measures, order statistics, and Kantorovich transport distances, Mem. Amer. Math. Soc. 261 (2019), no. 1259, v+126.
  • [BL2] by same author, A simple Fourier analytic proof of the AKT optimal matching theorem, arXiv:1909.06193 [math.PR], 2019.
  • [BMPR] M. Bruveris, P. W. Michor, A. Parusiński, and A. Rainer, Moser’s theorem on manifolds with corners, Proc. Amer. Math. Soc. 146 (2018), no. 11, 4889–4897.
  • [BottTu] R. Bott and L. W. Tu, Differential forms in algebraic topology, Graduate Texts in Mathematics, no. 82, Springer, 1982.
  • [EPC+] D. Epstein, M. Paterson, J. Cannon, D. Holt, S. Levy, and W. P. Thurston, Word processing in groups, Jones and Bartlett, 1992.
  • [E-Z] Ch. Even-Zohar, Models of random knots, J. Appl. Comput. Topol. 1 (2017), no. 2, 263–296.
  • [Fed] H. Federer, Geometric measure theory, Die Grundlehren der mathematischen Wissenschaften, Band 153, Springer-Verlag New York Inc., New York, 1969.
  • [FF] H. Federer and W. H. Fleming, Normal and integral currents, Ann. of Math. (2) 72 (1960), 458–520.
  • [FK] E. Flapan and K. Kozai, Linking number and writhe in random linear embeddings of graphs, J. Math. Chem. 54 (2016), no. 5, 1117–1133.
  • [FKW] P. Franek, M. Krčál, and H. Wagner, Solving equations and optimization problems with uncertainty, J. Appl. Comput. Topol. 1 (2018), no. 3-4, 297–330.
  • [GM] M. Gromov and V. D. Milman, A topological application of the isoperimetric inequality, American Journal of Mathematics 105 (1983), no. 4, 843–854.
  • [Gro] M. Gromov, Homotopical effects of dilatation, J. Differential Geom. 13 (1978), no. 3, 303–310.
  • [HPZ] N. Holden, Y. Peres, and A. Zhai, Gravitational allocation on the sphere, Proc. Natl. Acad. Sci. USA 115 (2018), no. 39, 9666–9671.
  • [HS] R. Hardt and L. Simon, Boundary regularity and embedded solutions for the oriented Plateau problem, Ann. of Math. (2) 110 (1979), no. 3, 439–486.
  • [Jones] P. W. Jones, Rectifiable sets and the traveling salesman problem, Invent. Math. 102 (1990), no. 1, 1–15.
  • [Led] M. Ledoux, The concentration of measure phenomenon, Mathematical Surveys and Monographs, vol. 89, American Mathematical Society, Providence, RI, 2001.
  • [Marko] J. F. Marko, Linking topology of tethered polymer rings with applications to chromosome segregation and estimation of the knotting length, Phys. Rev. E 79 (2009), 051905.
  • [Mil] K. C. Millett, Monte Carlo explorations of polygonal knot spaces, Knots in Hellas ’98 (Delphi), Ser. Knots Everything, vol. 24, World Sci. Publ., River Edge, NJ, 2000, pp. 306–334.
  • [Mor] F. Morgan, Geometric measure theory: A beginner’s guide, fourth ed., Elsevier/Academic Press, Amsterdam, 2009.
  • [Moser] J. Moser, On the volume elements on a manifold, Trans. Amer. Math. Soc. 120 (1965), 286–294.
  • [Shor] P. W. Shor, The average-case analysis of some on-line algorithms for bin packing, Combinatorica 6 (1986), no. 2, 179–200.
  • [Tal1] M. Talagrand, Matching random samples in many dimensions, Ann. Appl. Probab. 2 (1992), no. 4, 846–856.
  • [Tal2] by same author, The transportation cost from the uniform measure to the empirical measure in dimension 3\geq 3, Ann. Probab. 22 (1994), no. 2, 919–959.
  • [Tal3] by same author, Upper and lower bounds for stochastic processes: Modern methods and classical problems, Ergebnisse der Mathematik und ihrer Grenzgebiete. 3. Folge. A Series of Modern Surveys in Mathematics, vol. 60, Springer, Heidelberg, 2014.
  • [Tanaka] F. Tanaka, Gauge Theory of Topological Entanglements. I: — General Theory —, Progress of Theoretical Physics 68 (1982), no. 1, 148–163.
  • [Young] R. Young, Quantitative nonorientability of embedded cycles, Duke Math. J. 167 (2018), no. 1, 41–108.