Necklace splitting problem

Last updated
Example of necklace splitting with k = 2 (i.e. two partners), and t = 2 (i.e. two types of beads, here 8 red and 6 green). A 2-split is shown: one partner receives the largest section, and the other receives the remaining two pieces. Collier-de-perles-rouge-vert.svg
Example of necklace splitting with k = 2 (i.e. two partners), and t = 2 (i.e. two types of beads, here 8 red and 6 green). A 2-split is shown: one partner receives the largest section, and the other receives the remaining two pieces.

Necklace splitting is a picturesque name given to several related problems in combinatorics and measure theory. Its name and solutions are due to mathematicians Noga Alon [1] and Douglas B. West. [2]

Contents

The basic setting involves a necklace with beads of different colors. The necklace should be divided between several partners (e.g. thieves), such that each partner receives the same amount of every color. Moreover, the number of cuts should be as small as possible (in order to waste as little as possible of the metal in the links between the beads).

Variants

The following variants of the problem have been solved in the original paper:

  1. Discrete splitting: [1] :Th 1.1 The necklace has beads. The beads come in different colors. There are beads of each color , where is a positive integer. Partition the necklace into parts (not necessarily contiguous), each of which has exactly beads of color i. Use at most cuts. Note that if the beads of each color are contiguous on the necklace, then at least cuts must be done inside each color, so is optimal.
  2. Continuous splitting: [1] :Th 2.1 The necklace is the real interval . Each point of the interval is colored in one of different colors. For every color , the set of points colored by is Lebesgue-measurable and has length , where is a non-negative real number. Partition the interval to parts (not necessarily contiguous), such that in each part, the total length of color is exactly . Use at most cuts.
  3. Measure splitting: [1] :Th 1.2 The necklace is a real interval. There are different measures on the interval, all absolutely continuous with respect to length. The measure of the entire necklace, according to measure , is . Partition the interval to parts (not necessarily contiguous), such that the measure of each part, according to measure , is exactly . Use at most cuts. This is a generalization of the Hobby–Rice theorem, and it is used to get an exact division of a cake.

Each problem can be solved by the next problem:

Proof

The case can be proved by the Borsuk-Ulam theorem. [2]

When is an odd prime number, the proof involves a generalization of the Borsuk-Ulam theorem. [3]

When is a composite number, the proof is as follows (demonstrated for the measure-splitting variant). Suppose . There are measures, each of which values the entire necklace as . Using cuts, divide the necklace to parts such that measure of each part is exactly . Using cuts, divide each part to parts such that measure of each part is exactly . All in all, there are now parts such that measure of each part is exactly . The total number of cuts is plus which is exactly .

Further results

Splitting random necklaces

In some cases, random necklaces can be split equally using fewer cuts. Mathematicians Noga Alon, Dor Elboim, Gábor Tardos and János Pach studied the typical number of cuts required to split a random necklace between two thieves. [4] In the model they considered, a necklace is chosen uniformly at random from the set of necklaces with t colors and m beads of each color. As m tends to infinity, the probability that the necklace can be split using ⌊(t + 1)/2⌋ cuts or less tends to zero while the probability that it's possible to split with ⌊(t + 1)/2⌋ + 1 cuts is bounded away from zero. More precisely, letting X = X(t,m) be the minimal number of cuts required to split the necklace. The following holds as m tends to infinity. For any

For any

Finally, when is odd and

One can also consider the case in which the number of colors tends to infinity. When m=1 and the t tends to infinity, the number of cuts required is at most 0.4t and at least 0.22t with high probability. It is conjectured that there exists some 0.22 < c < 0.4 such that X(t,1)/t  converges to c in distribution.

One cut fewer than needed

In the case of two thieves [i.e. k = 2] and t colours, a fair split would require at most t cuts. If, however, only t  1 cuts are available, Hungarian mathematician Gábor Simonyi [5] shows that the two thieves can achieve an almost fair division in the following sense.

If the necklace is arranged so that no t-split is possible, then for any two subsets D1 and D2 of { 1, 2, ...,  t }, not both empty, such that , a (t  1)-split exists such that:

This means that if the thieves have preferences in the form of two "preference" sets D1 and D2, not both empty, there exists a (t  1)-split so that thief 1 gets more beads of types in his preference set D1 than thief 2; thief 2 gets more beads of types in her preference set D2 than thief 1; and the rest are equal.

Simonyi credits Gábor Tardos with noticing that the result above is a direct generalization of Alon's original necklace theorem in the case k = 2. Either the necklace has a (t  1)-split, or it does not. If it does, there is nothing to prove. If it does not, we may add beads of a fictitious colour to the necklace, and make D1 consist of the fictitious colour and D2 empty. Then Simonyi's result shows that there is a t-split with equal numbers of each real colour.

Negative result

For every there is a measurable -coloring of the real line such that no interval can be fairly split using at most cuts. [6]

Splitting multidimensional necklaces

The result can be generalized to n probability measures defined on a d dimensional cube with any combination of n(k  1) hyperplanes parallel to the sides for k thieves. [7]

Approximation algorithm

An approximation algorithm for splitting a necklace can be derived from an algorithm for consensus halving. [8]

See also

Related Research Articles

In mathematics, the Cantor set is a set of points lying on a single line segment that has a number of unintuitive properties. It was discovered in 1874 by Henry John Stephen Smith and introduced by German mathematician Georg Cantor in 1883.

In graph theory, an expander graph is a sparse graph that has strong connectivity properties, quantified using vertex, edge or spectral expansion. Expander constructions have spawned research in pure and applied mathematics, with several applications to complexity theory, design of robust computer networks, and the theory of error-correcting codes.

In measure theory, a branch of mathematics, the Lebesgue measure, named after French mathematician Henri Lebesgue, is the standard way of assigning a measure to subsets of n-dimensional Euclidean space. For n = 1, 2, or 3, it coincides with the standard measure of length, area, or volume. In general, it is also called n-dimensional volume, n-volume, or simply volume. It is used throughout real analysis, in particular to define Lebesgue integration. Sets that can be assigned a Lebesgue measure are called Lebesgue-measurable; the measure of the Lebesgue-measurable set A is here denoted by λ(A).

<span class="mw-page-title-main">Riemann integral</span> Basic integral in elementary calculus

In the branch of mathematics known as real analysis, the Riemann integral, created by Bernhard Riemann, was the first rigorous definition of the integral of a function on an interval. It was presented to the faculty at the University of Göttingen in 1854, but not published in a journal until 1868. For many functions and practical applications, the Riemann integral can be evaluated by the fundamental theorem of calculus or approximated by numerical integration, or simulated using Monte Carlo Integration.

In vector calculus and differential geometry the generalized Stokes theorem, also called the Stokes–Cartan theorem, is a statement about the integration of differential forms on manifolds, which both simplifies and generalizes several theorems from vector calculus. In particular, the fundamental theorem of calculus is the special case where the manifold is a line segment, Green’s theorem and Stokes' theorem are the cases of a surface in or and the divergence theorem is the case of a volume in Hence, the theorem is sometimes referred to as the Fundamental Theorem of Multivariate Calculus.

Van der Waerden's theorem is a theorem in the branch of mathematics called Ramsey theory. Van der Waerden's theorem states that for any given positive integers r and k, there is some number N such that if the integers {1, 2, ..., N} are colored, each with one of r different colors, then there are at least k integers in arithmetic progression whose elements are of the same color. The least such N is the Van der Waerden number W(rk), named after the Dutch mathematician B. L. van der Waerden.

The necklace problem is a problem in recreational mathematics concerning the reconstruction of necklaces from partial information.

In mathematics, a measure-preserving dynamical system is an object of study in the abstract formulation of dynamical systems, and ergodic theory in particular. Measure-preserving systems obey the Poincaré recurrence theorem, and are a special case of conservative systems. They provide the formal, mathematical basis for a broad range of physical systems, and, in particular, many systems from classical mechanics as well as systems in thermodynamic equilibrium.

Chebotarev's density theorem in algebraic number theory describes statistically the splitting of primes in a given Galois extension K of the field of rational numbers. Generally speaking, a prime integer will factor into several ideal primes in the ring of algebraic integers of K. There are only finitely many patterns of splitting that may occur. Although the full description of the splitting of every prime p in a general Galois extension is a major unsolved problem, the Chebotarev density theorem says that the frequency of the occurrence of a given pattern, for all primes p less than a large integer N, tends to a certain limit as N goes to infinity. It was proved by Nikolai Chebotaryov in his thesis in 1922, published in.

In mathematics, the total variation identifies several slightly different concepts, related to the (local or global) structure of the codomain of a function or a measure. For a real-valued continuous function f, defined on an interval [a, b] ⊂ R, its total variation on the interval of definition is a measure of the one-dimensional arclength of the curve with parametric equation xf(x), for x ∈ [a, b]. Functions whose total variation is finite are called functions of bounded variation.

In mathematical measure theory, for every positive integer n the ham sandwich theorem states that given n measurable "objects" in n-dimensional Euclidean space, it is possible to divide each one of them in half (with respect to their measure, e.g. volume) with a single (n − 1)-dimensional hyperplane. This is even possible if the objects overlap.

<span class="mw-page-title-main">Szemerédi regularity lemma</span>

Szemerédi's regularity lemma is one of the most powerful tools in extremal graph theory, particularly in the study of large dense graphs. It states that the vertices of every large enough graph can be partitioned into a bounded number of parts so that the edges between different parts behave almost randomly.

An envy-free cake-cutting is a kind of fair cake-cutting. It is a division of a heterogeneous resource ("cake") that satisfies the envy-free criterion, namely, that every partner feels that their allocated share is at least as good as any other share, according to their own subjective valuation.

<span class="mw-page-title-main">Necklace (combinatorics)</span>

In combinatorics, a k-ary necklace of length n is an equivalence class of n-character strings over an alphabet of size k, taking all rotations as equivalent. It represents a structure with n circularly connected beads which have k available colors.

In additive number theory and combinatorics, a restricted sumset has the form

The mathematical discipline of topological combinatorics is the application of topological and algebro-topological methods to solving problems in combinatorics.

Exact division, also called consensus division, is a partition of a continuous resource ("cake") into some k pieces, such that each of n people with different tastes agree on the value of each of the pieces. For example, consider a cake which is half chocolate and half vanilla. Alice values only the chocolate and George values only the vanilla. The cake is divided into three pieces: one piece contains 20% of the chocolate and 20% of the vanilla, the second contains 50% of the chocolate and 50% of the vanilla, and the third contains the rest of the cake. This is an exact division (with k = 3 and n = 2), as both Alice and George value the three pieces as 20%, 50% and 30% respectively. Several common variants and special cases are known by different terms:

In mathematics, and in particular the necklace splitting problem, the Hobby–Rice theorem is a result that is useful in establishing the existence of certain solutions. It was proved in 1965 by Charles R. Hobby and John R. Rice; a simplified proof was given in 1976 by A. Pinkus.

In mathematics, a line integral is an integral where the function to be integrated is evaluated along a curve. The terms path integral, curve integral, and curvilinear integral are also used; contour integral is used as well, although that is typically reserved for line integrals in the complex plane.

In the fair cake-cutting problem, the partners often have different entitlements. For example, the resource may belong to two shareholders such that Alice holds 8/13 and George holds 5/13. This leads to the criterion of weighted proportionality (WPR): there are several weights that sum up to 1, and every partner should receive at least a fraction of the resource by their own valuation.

References

  1. 1 2 3 4 5 6 Alon, Noga (1987). "Splitting Necklaces". Advances in Mathematics . 63 (3): 247–253. doi: 10.1016/0001-8708(87)90055-7 .
  2. 1 2 Alon, Noga; West, Douglas B. (December 1986). "The Borsuk-Ulam theorem and bisection of necklaces". Proceedings of the American Mathematical Society . 98 (4): 623–628. doi: 10.1090/s0002-9939-1986-0861764-9 .
  3. I.Barany and S.B.Shlosman and A.Szucs (1981). "On a topological generalization of a theorem of Tverberg". Journal of the London Mathematical Society. 2 (23): 158–164. CiteSeerX   10.1.1.640.1540 . doi:10.1112/jlms/s2-23.1.158.
  4. Alon, Noga; Elboim, Dor; Tardos, Gábor; Pach, János (2021). "Random Necklaces require fewer cuts". arXiv: 2112.14488 [math.CO].
  5. Simonyi, Gábor (2008). "Necklace bisection with one cut less than needed". Electronic Journal of Combinatorics. 15 (N16). doi: 10.37236/891 .
  6. Alon, Noga (November 25, 2008). "Splitting necklaces and measurable colorings of the real line". Proceedings of the American Mathematical Society . 137 (5): 1593–1599. arXiv: 1412.7996 . doi: 10.1090/s0002-9939-08-09699-8 . ISSN   1088-6826.
  7. de Longueville, Mark; Rade T. Živaljević (2008). "Splitting multidimensional necklaces". Advances in Mathematics . 218 (3): 926–939. arXiv: math/0610800 . doi: 10.1016/j.aim.2008.02.003 .
  8. Simmons, Forest W.; Su, Francis Edward (February 2003). "Consensus-halving via theorems of Borsuk-Ulam and Tucker". Mathematical Social Sciences. 45 (1): 15–25. CiteSeerX   10.1.1.203.1189 . doi:10.1016/s0165-4896(02)00087-2.