HOW TO LOOK AT MINKOWSKI S THEOREM

Similar documents
GAUSS CIRCLE PROBLEM

DECOUPLING LECTURE 6

Lebesgue Measure on R n

Large Deviations for Weakly Dependent Sequences: The Gärtner-Ellis Theorem

Connectedness. Proposition 2.2. The following are equivalent for a topological space (X, T ).

ON SPACE-FILLING CURVES AND THE HAHN-MAZURKIEWICZ THEOREM

approximation of the dimension. Therefore, we get the following corollary.

Additive Combinatorics Lecture 12

MAT2342 : Introduction to Applied Linear Algebra Mike Newman, fall Projections. introduction

Introduction to Real Analysis Alternative Chapter 1

After taking the square and expanding, we get x + y 2 = (x + y) (x + y) = x 2 + 2x y + y 2, inequality in analysis, we obtain.

A VERY BRIEF REVIEW OF MEASURE THEORY

Course Description for Real Analysis, Math 156

We set up the basic model of two-sided, one-to-one matching

Commutative Banach algebras 79

Decoupling course outline Decoupling theory is a recent development in Fourier analysis with applications in partial differential equations and

GAUSS S CIRCLE PROBLEM

Linear Algebra, Summer 2011, pt. 2

Modern Algebra Prof. Manindra Agrawal Department of Computer Science and Engineering Indian Institute of Technology, Kanpur

DR.RUPNATHJI( DR.RUPAK NATH )

THE FINITE FIELD KAKEYA CONJECTURE

1.5 Approximate Identities

Fourier Transform & Sobolev Spaces

SPECIAL CASES OF THE CLASS NUMBER FORMULA

Lecture 4: Constructing the Integers, Rationals and Reals

Lebesgue Measure on R n

Math 52: Course Summary

MITOCW watch?v=rf5sefhttwo

NOTES ON DIFFERENTIAL FORMS. PART 1: FORMS ON R n

Chapter 1 Review of Equations and Inequalities

Metric spaces and metrizability

SPECIAL POINTS AND LINES OF ALGEBRAIC SURFACES

Lecture for Week 2 (Secs. 1.3 and ) Functions and Limits

Existence and Uniqueness

Chapter 11 - Sequences and Series

The Banach-Tarski paradox

Spanning, linear dependence, dimension

Linear Algebra. Preliminary Lecture Notes

D(f/g)(P ) = D(f)(P )g(p ) f(p )D(g)(P ). g 2 (P )

AN INTEGRAL FORMULA FOR TRIPLE LINKING IN HYPERBOLIC SPACE

2 Metric Spaces Definitions Exotic Examples... 3

We have to prove now that (3.38) defines an orthonormal wavelet. It belongs to W 0 by Lemma and (3.55) with j = 1. We can write any f W 1 as

EXPOSITORY NOTES ON DISTRIBUTION THEORY, FALL 2018

Solutions to Problem Set 5 for , Fall 2007

Instructor (Brad Osgood)

Spanning and Independence Properties of Finite Frames

Overview of normed linear spaces

LINEAR ALGEBRA: THEORY. Version: August 12,

Linear Algebra. Preliminary Lecture Notes

MITOCW ocw f99-lec01_300k

Quantum Mechanics-I Prof. Dr. S. Lakshmi Bala Department of Physics Indian Institute of Technology, Madras. Lecture - 21 Square-Integrable Functions

Calculus of One Real Variable Prof. Joydeep Dutta Department of Economic Sciences Indian Institute of Technology, Kanpur

LECTURE 10: REVIEW OF POWER SERIES. 1. Motivation

Error Correcting Codes Prof. Dr. P. Vijay Kumar Department of Electrical Communication Engineering Indian Institute of Science, Bangalore

COMPLEX ANALYSIS Spring 2014

Chapter Four Gelfond s Solution of Hilbert s Seventh Problem (Revised January 2, 2011)

Cosets and Lagrange s theorem

Hausdorff Measure. Jimmy Briggs and Tim Tyree. December 3, 2016

chapter 12 MORE MATRIX ALGEBRA 12.1 Systems of Linear Equations GOALS

CHAPTER 1. Introduction

Central limit theorem. Paninski, Intro. Math. Stats., October 5, probability, Z N P Z, if

7.2 Conformal mappings

GAUSSIAN ELIMINATION AND LU DECOMPOSITION (SUPPLEMENT FOR MA511)

An Intuitive Introduction to Motivic Homotopy Theory Vladimir Voevodsky

7.1 Indefinite Integrals Calculus

Comparing continuous and discrete versions of Hilbert s thirteenth problem

Math 320-3: Lecture Notes Northwestern University, Spring 2015

Both these computations follow immediately (and trivially) from the definitions. Finally, observe that if f L (R n ) then we have that.

Sequence convergence, the weak T-axioms, and first countability

Lebesgue-Radon-Nikodym Theorem

S chauder Theory. x 2. = log( x 1 + x 2 ) + 1 ( x 1 + x 2 ) 2. ( 5) x 1 + x 2 x 1 + x 2. 2 = 2 x 1. x 1 x 2. 1 x 1.

Math 676. A compactness theorem for the idele group. and by the product formula it lies in the kernel (A K )1 of the continuous idelic norm

ON THE PROBLEM OF CHARACTERIZING DERIVATIVES

Topological properties of Z p and Q p and Euclidean models

We are going to discuss what it means for a sequence to converge in three stages: First, we define what it means for a sequence to converge to zero

Sphere Packings, Coverings and Lattices

Slope Fields: Graphing Solutions Without the Solutions

MATH 411 NOTES (UNDER CONSTRUCTION)

Spectral Graph Theory Lecture 2. The Laplacian. Daniel A. Spielman September 4, x T M x. ψ i = arg min

_CH04_p pdf Page 52

Chapter 6: The Definite Integral

Topological properties

ACCESS TO SCIENCE, ENGINEERING AND AGRICULTURE: MATHEMATICS 1 MATH00030 SEMESTER / Lines and Their Equations

Regularity for Poisson Equation

Final Review Sheet. B = (1, 1 + 3x, 1 + x 2 ) then 2 + 3x + 6x 2

MA554 Assessment 1 Cosets and Lagrange s theorem

Algebraic geometry of the ring of continuous functions

Uniformly discrete forests with poor visibility

Mathematical Research Letters 10, (2003) THE FUGLEDE SPECTRAL CONJECTURE HOLDS FOR CONVEX PLANAR DOMAINS

Introductory Analysis I Fall 2014 Homework #5 Solutions

0. Introduction 1 0. INTRODUCTION

The principle of concentration-compactness and an application.

The uniformization theorem

SYDE 112, LECTURE 7: Integration by Parts

Ω Ω /ω. To these, one wants to add a fourth condition that arises from physics, what is known as the anomaly cancellation, namely that

Sequences and Series of Functions

UMASS AMHERST MATH 300 SP 05, F. HAJIR HOMEWORK 8: (EQUIVALENCE) RELATIONS AND PARTITIONS

Transform methods. and its inverse can be used to analyze certain time-dependent PDEs. f(x) sin(sxπ/(n + 1))

Abstract & Applied Linear Algebra (Chapters 1-2) James A. Bernhard University of Puget Sound

Lecture 2: Vector Spaces, Metric Spaces

Transcription:

HOW TO LOOK AT MINKOWSKI S THEOREM COSMIN POHOATA Abstract. In this paper we will discuss a few ways of thinking about Minkowski s lattice point theorem. The Minkowski s Lattice Point Theorem essentially states that bounded, convex, centrally symmetric sets in the n-dimensional Euclidean space R n will always contain a nontrivial lattice point of Z n, provided that their volume is big enough. This simple and natural result was proved by Hermann Minkowski in 889 and became the foundation of the so-called geometry of numbers, a field which connects multiple important areas of mathematics, such as algebraic number theory, harmonic analysis or complexity theory. Minkowski s Theorem also appears in discrete/olympiad mathematics and can prove to be quite an important tool when dealing with diophantine equations or density results. We refer to [] for a collection of applications of this kind. In this paper we will not dwell on the applications, but only present a few different ways of looking at Minkowski s theorem, with the hope of sheding some light on its broad ramifications. We begin by stating a more precise version of Minkowski s theorem. Theorem. Let K R n be a bounded, convex, centrally symmetric set. If in addition the volume of K satisfies vol K > n, then K contains at least one non-trivial lattice point of Z n. Let us first clarify a little what all these terms mean and try to make a little more sense about the statement. A subset K of R n is bounded if it is contained in a ball of sufficiently large radius. Such a set K is furthermore called convex if for any two points x, y in K, the line segment {tx + ( ty, 0 t } is also in K and centrally symmetric if the origin we fix for the space R n is in K and x is in K if and only if x is in K. Now, let us also note that the n bound is tight. Indeed, pick the cube Q = (, n R n. This indeed has volume n, it is bounded, convex, and centrally symmetric - but it doesn t contain any other lattice point of Z n apart from the origin. What s special about this counter-example? Well, it doesn t contain its boundary! And the following result tells us that this is what actually differentiates it from all the good sets. Theorem. Let K R n be a bounded, convex, centrally symmetric set, which in addition is also compact (thus contains its boundary. If the volume of K satisfies vol K n, then K contains at least one non-trivial lattice point of Z n. Fortunately, Theorem is nothing but a simple consequence of Theorem and an elementary compactness argument. Indeed, suppose that the volume of K satisfies vol K = n (since otherwise, Theorem follows from Theorem even without the compactness property of K. For each ɛ > 0, let K ɛ be the dilate K( + ɛ. Notice that the sets K ɛ satisfy the assumption that vol K ɛ > n, thus by Theorem, K ɛ contains a nonzero lattice point of Z n. But K is bounded, so there are only finitely many possiblities for this

COSMIN POHOATA nonzero lattice point for each ɛ. Thus, we can find a sequence of ɛ s tending to 0 for which this lattice point is the same. The convexity of the sets K ɛ, in combination with the fact that the sets contain 0, implies that the sets are nested, and therefore this lattice point lies in K ɛ for all ɛ > 0. Since K is compact, we have that K = ɛ>0 K ɛ, and therefore this lattice point lies in K. Consequently, in this paper we will just focus on Theorem and talk about bounded, convex, centrally-symmetric sets K that are not necessarily compact. In addition, we should add that Theorem can actually be stated in more general terms, where, for example, Z n is replaced by any lattice L inside R n and where the condition on the volume is modified to vol K > n det L, where det L denotes the volume of the lattice s fundamental parallelipiped. As a matter of fact, this is the version which Minkowski really proved and what circulates nowadays in literature as Minkowski s theorem. However, for the purposes of this exposition, we prefer to work with the simpler (to visualize Z n lattice and so we will use the statement cited above. Accordingly, whenever we refer to a point as being a lattice point, we mean a point from the integer lattice. The first way of thinking about Minkowski s theorem, even for the analysis-inclined people, would still be to draw a picture for the two-dimensional case and see if any property of the Euclidean plane would give the existence of the non-zero lattice point inside K. Indeed, this works! First Proof of Theorem. The idea is to look again at the cube Q = [, ] n (this time its closed version. Note that this cube is centered at the origin and that all its translates by even coordinate vectors partition R n. More formally, we can thus say that R n = (Q + u. u Z n For conveniece, let us denote Q + u by Q u. Note that K is bounded, thus K intersects only a finite set of these Q u s, call it Q. Now, let us look at the sets Q u from Q and their translations back to Q. These translations will create an agglomeration of parts of K inside Q. However, we know that vol K > n, whereas vol Q = n. Therefore, there will be at least an overlap of two translated Q u s; pick some point x lying in this overlap. This point x can be thus written as x = v + y = w + z for some distinct points y, z in K and some distinct vectors v, w in Z n. In particular, we get that the point y z = w v is in Z n. But y K and z K (as z is in K and K is centrally symmetric; thus the convexity of K yields that y z is also K, which means that y z is a non-zero lattice point that lies in K. This proves Minkowski s theorem. This first proof is perhaps the most well-known as it appears in almost all textbooks that include some geometry of numbers. The argument is most commonly attributed to H. E. Blichfeld or to Minkowski himself. In fact, the version for general lattices L only requires us to map the lattice to Z n via the natural linear transformation taking the (supposedly n linearly independent vectors defining L to the standard basis vectors e i = (0,...,,..., 0 of R n. This will immediately reduce everything to Theorem. However, some of the proofs that we will give will require some extra-care about how this linear map works, so we chose the simpler version in order to avoid all these technical details.

HOW TO LOOK AT MINKOWSKI S THEOREM 3 The second (incomplete proof turns out to be more of an heuristic argument where we use an apparently completely different idea involving Fourier analysis. However, this will lead us to a very short proof that uses only a simple integration trick. Second Proof of Theorem. The key here is to begin by noticing that the number of lattice points inside K is given by the cardinality of K Z n = u Z n χ K (u, where χ K represents the characteristic function of K. Then, what we do is seek to apply the Poisson Summation Formula 3. Let f together with its Fourier transform f be continuous functions on R n that have moderate decrease, or in other words satisfy f(x A + x n+ɛ and f(x B + x n+ɛ for some positive constants A, B, and some small sufficiently small ɛ > 0. Then, f(x + u = f(ue πixu. u Z n u Z n In particular, if we take x = 0, this gives us f(u = f(u. u Z n u Z n Already seeing a resemblance with the right hand side of our initial identity, we would like to write that K Z n = χ K (u = χ K (u. u Z n u Z n But we are not allowed to do this, since the characteristic function χ K is far from being a continuous function, let alone a continuous function of moderate decrease. However, if we could find some continuous function of moderate decrease f : R n R which satisfies χ K f, then this problem would be fixed! We would then be able to write down that K Z n = χ K (u f(u = f(u. u Z n u Z n u Z n Moreover, if we would also be able to cook up this nice function f to have the additional property that its Fourier transform f is positive real-valued, then we would get K Z n f(0 = f(xdx, R n and this would gives us a nice lower bound (provided that f is simple enough. In general, the characteristic function χ K of a subset S of some bigger set X is defined to be the function χ K : X {0, } so that χ K (x = if x S and χ K (x = 0 otherwise. 3 Again, this happens to be a theorem which admits a lot of different formulations depending on the level of generality one wants to use. We refer to [] and [3] for various discussions of these aspects.

4 COSMIN POHOATA Now, convolutions come to mind! Indeed, given two integrable, bounded complexvalued functions F and G on R n, it is a known fact from basic Fourier analysis that their convolution (F G(x = F (yg(x ydy = R n F (x yg(ydy. R n is continuous. We refer to [] for a proof. Hence, a function ( such as f = F G is what we are looking for. For example, for the function f = χ vol K K χ K, we also get immediately that ( f(x = vol K χ K χ K (x = vol K χ K (yχ K (x ydy R n vol K χ K (ydy R n vol K vol K = χ K (x for all x in K, while for x K, we have that f(x = vol K χ K (yχ K (x ydy = 0. R n (since y and x y can never both be in K we have that χ K f. Moreover, f = ( vol K χ K χ K = vol K as long as x is not in K - by convexity. Thus, ( χ K χ K = vol K χ K > 0. Thus f is almost everything we wished for: it is f, satisfies χ K f and f is positive real-valued. Unfortunately, it is not of moderate decrease (even though it comes pretty close. So, tehnically, we cannot apply the Poisson Summation Formula to it. But note that if we could have applied it, then we would have been done. Indeed, writing K Z n f(0 = vol K χ K (0 = vol K χ K (xdx R n = vol K = vol K n >, gives us that K contains at least one non-trivial lattice point. This would work if we would have some additional assumptions on the boundary of K. For example, if it were for K to have non-vanishing Gauss curvature at each point, then Corollary 3.3 of [3, pp. 336] tells us that χ K (x = O ( x n+, as x ; thus f(x = vol K χ K = O ( x n, as x, and so we would have moderate decrease. But let s not complicate things unnecessarily. How could we get a function that looks like f but has moderate decrease? In [3], this is done by introducing a so-called bump-function (or approximation to the identity φ

HOW TO LOOK AT MINKOWSKI S THEOREM 5 which is non-negative, smooth, supported in the unit ball and satisfying φ(xdx =, and then taking the convolution f = χ K φ (modulo some rescaling. This fixes everything, since now the convolution represents a smooth function of compact support, so we can actually apply the Poisson Summation formula to it. While very natural, this idea however complicates the computations, since it doesn t give us the nice right-hand side vol K n we got above, and we have to perform some other several tricks on the left-hand side in order to get what we want. We will not proceed this further here. We refer to [3] for the details of computation, where the authors deal with the estimation of the number of lattice points in higher dimensional convex bodies. What we will do instead is pick-up this analysis-oriented idea and give it a different turn, without involving the Poisson Summation Formula. Third Proof of Theorem. Consider the function Ψ : R n R defined by Ψ(x = (x + u. u Z n χ K As advertised, this should resemble the first line of the previous proof, since Φ(x is nothing but the cardinality of the intersection K (x + Z n. Now, instead of seeking to apply the Poisson Summation Formula to this expression of Ψ(x, we integrate it over the closed cube [0, ] n R n. We can obviously do this, since Ψ(x is bounded (as K is bounded and integrable (as it is a sum of integrable functions. Thus, we can write [0,] n Ψ(xdx = ( χ K (x + u dx [0,] n u Z n = ( χ K (x + udx u Z n [0,] n = ( = u Z n R n χ K = vol K = vol K n >. χ K (xdx u+[0,] n (xdx since R n = by changing the order of summation by doing a change of variables u Z n (u + [0, ] n It follows that there must be a point x in [0, ] n so that Ψ(x. Indeed, if this weren t the case, then, since Ψ(x is integer valued, we would have Ψ(x, and so we would get Ψ(xdx >, [0,] n which is obviously impossible. So, there must be at least two points u, u in Z n so that x + u, x + u are both in K. Call these two points y, z, with y z both in K. Then, the point u u = y z is in K (by central symmetry and convexity; but it is also in Z n because both u and u are in Z n ; hence we get that y z is a nonzero lattice point in K, which proves Minkowski once again.

6 COSMIN POHOATA Notice the fact that this third proof is a formalization of what happened in the previous one when we took the translates of the cubes that intersected the set K. Also, the computation above might actually be the proof of the Poisson Summation Formula in disguise! There s also one fourth way to look at Minkowski s theorem. After all, the result is a statement about the lattice point contained inside a set K, so it should have some connection with the celebrated Gauss Circle Problem. How many lattice points lie inside a circle in R centered at the origin and having radius r > 0? The area inside a circle of radius r is given by πr, and since a square of area in R contains one integer point, the expected answer to the problem could be about πr. In fact, it should be slightly higher than this, since circles are more efficient at enclosing space than squares. Therefore in fact it should be expected that N(r = πr + E(r, for some error term E(r. Finding an upper bound for this error term is practically what this problem has turned into. Gauss himself proved that E(r πr with a very simple argument which can be found in [4], while conjecturing that E(r = O(r +ɛ. This is still open today, with the best known bound being E(r = O(r 3. We will refer again to [3] for the proof of this fact (which goes along the line of our second proof of Minkowski s theorem. Anyway, what we will need here for our purposes is the result that if M(r denotes the number of r -lattice points of the bounded, convex, centrally symmetric set K, i.e. points x K so that rx Z n, then M(r lim r r n = vol K. This comes very natural after the Gauss circle problems estimates (which generalize to higher dimensions. Another way to prove this would be to appeal to the following more general Lemma about Jordan measurable sets (corroborated with the fact that every bounded convex set is Jordan measurable - see [5]. Lemma. Let K R n be a (bounded Jordan measurable set. Then, using the notations from above, M(r lim r r n = vol K. Sketch of Proof. By scalling appropriately, we may assume that K (, n. Then, note that for any positive integer r, the fraction M(r r is precisely a Riemann sum for the n characteristic function χ K corresponding to the partition of [, ] n into subsquares of sidelength r, so the convergence of these sums to χ K = vol K as r is immediate from the definition of Jordan measurability.

HOW TO LOOK AT MINKOWSKI S THEOREM 7 Fourth Proof of Theorem. Let us now work with K instead of K, and so denote by M(r the number of r -lattice points of K. Clearly, K is bounded and convex, so it is Jordan measurable. Hence, M(r lim r r n = vol K = vol K n >, which means that there is some r 0 such that for each r > r 0, we have M(r r n >, i.e. M(r > r n = (Z/rZ n. Hence, for these big r s, by the pidgeonhole principle, there exist two distinct r -lattice points x, y of K so that rx and ry have the same coordinates modulo r. This means that x y = rx ry r is in Z n. However, x, y are in K so x y is in K too (by central symmetry and convexity. Hence, we get a nonzero lattice point x y in K, thus proving Minkowski s theorem. References [] T. Andreescu, G. Dospinescu, Problems from the Book, XYZ Press, 00x. [] E. M. Stein, R. Shakarchi, Fourier Analysis: An Introduction, Princeton University Press, 003. [3] E. M. Stein, R. Shakarchi, Functional Analysis: Introduction to Further Topics in Analysis, Princeton University Press, 0. [4] D. Hilbert, S. Cohn-Vossen, Geometry and the Imagination, New York: Chelsea, pp. 33-35, 999. [5] A. Frohlich, M. J. Taylor, Algebraic Number Theory, Cambridge University Press, 99. E-mail address: apohoata@princeton.edu