Dickson's lemma

Last updated

In mathematics, Dickson's lemma states that every set of -tuples of natural numbers has finitely many minimal elements. This simple fact from combinatorics has become attributed to the American algebraist L. E. Dickson, who used it to prove a result in number theory about perfect numbers. [1] However, the lemma was certainly known earlier, for example to Paul Gordan in his research on invariant theory. [2]

Contents

Example

Infinitely many minimal pairs of real numbers x,y (the black hyperbola) but only five minimal pairs of positive integers (red) have xy >= 9. Dickson hyperbola9.svg
Infinitely many minimal pairs of real numbers x,y (the black hyperbola) but only five minimal pairs of positive integers (red) have xy  9.

Let be a fixed natural number, and let be the set of pairs of numbers whose product is at least . When defined over the positive real numbers, has infinitely many minimal elements of the form , one for each positive number ; this set of points forms one of the branches of a hyperbola. The pairs on this hyperbola are minimal, because it is not possible for a different pair that belongs to to be less than or equal to in both of its coordinates. However, Dickson's lemma concerns only tuples of natural numbers, and over the natural numbers there are only finitely many minimal pairs. Every minimal pair of natural numbers has and , for if x were greater than K then (x 1, y) would also belong to S, contradicting the minimality of (x, y), and symmetrically if y were greater than K then (x, y 1) would also belong to S. Therefore, over the natural numbers, has at most minimal elements, a finite number. [note 1]

Formal statement

Let be the set of non-negative integers (natural numbers), let n be any fixed constant, and let be the set of -tuples of natural numbers. These tuples may be given a pointwise partial order, the product order, in which if and only if for every . The set of tuples that are greater than or equal to some particular tuple forms a positive orthant with its apex at the given tuple.

With this notation, Dickson's lemma may be stated in several equivalent forms:

Generalizations and applications

Dickson used his lemma to prove that, for any given number , there can exist only a finite number of odd perfect numbers that have at most prime factors. [1] However, it remains open whether there exist any odd perfect numbers at all.

The divisibility relation among the P-smooth numbers, natural numbers whose prime factors all belong to the finite set P, gives these numbers the structure of a partially ordered set isomorphic to . Thus, for any set S of P-smooth numbers, there is a finite subset of S such that every element of S is divisible by one of the numbers in this subset. This fact has been used, for instance, to show that there exists an algorithm for classifying the winning and losing moves from the initial position in the game of Sylver coinage, even though the algorithm itself remains unknown. [6]

The tuples in correspond one-for-one with the monomials over a set of variables . Under this correspondence, Dickson's lemma may be seen as a special case of Hilbert's basis theorem stating that every polynomial ideal has a finite basis, for the ideals generated by monomials. Indeed, Paul Gordan used this restatement of Dickson's lemma in 1899 as part of a proof of Hilbert's basis theorem. [2]

See also

Notes

  1. With more care, it is possible to show that one of and is at most , and that there is at most one minimal pair for each choice of one of the coordinates, from which it follows that there are at most minimal elements.

Related Research Articles

In mathematics, a set is countable if either it is finite or it can be made in one to one correspondence with the set of natural numbers. Equivalently, a set is countable if there exists an injective function from it into the natural numbers; this means that each element in the set may be associated to a unique natural number, or that the elements of the set can be counted one at a time, although the counting may never finish due to an infinite number of elements.

In mathematics, more specifically in general topology and related branches, a net or Moore–Smith sequence is a function whose domain is a directed set. The codomain of this function is usually some topological space. Nets directly generalize the concept of a sequence in a metric space. Nets are primarily used in the fields of Analysis and Topology, where they are used to characterize many important topological properties that, sequences are unable to characterize. Nets are in one-to-one correspondence with filters.

In mathematics, an ordered field is a field together with a total ordering of its elements that is compatible with the field operations. Basic examples of ordered fields are the rational numbers and the real numbers, both with their standard orderings.

<span class="mw-page-title-main">Partially ordered set</span> Mathematical set with an ordering

In mathematics, especially order theory, a partial order on a set is an arrangement such that, for certain pairs of elements, one precedes the other. The word partial is used to indicate that not every pair of elements needs to be comparable; that is, there may be pairs for which neither element precedes the other. Partial orders thus generalize total orders, in which every pair is comparable.

In mathematics, the branch of real analysis studies the behavior of real numbers, sequences and series of real numbers, and real functions. Some particular properties of real-valued sequences and functions that real analysis studies include convergence, limits, continuity, smoothness, differentiability and integrability.

In mathematics, a well-order on a set S is a total ordering on S with the property that every non-empty subset of S has a least element in this ordering. The set S together with the ordering is then called a well-ordered set. In some academic articles and textbooks these terms are instead written as wellorder, wellordered, and wellordering or well order, well ordered, and well ordering.

In mathematics, the infimum of a subset of a partially ordered set is the greatest element in that is less than or equal to each element of if such an element exists. In other words, it is the greatest element of that is lower or equal to the lowest element of . Consequently, the term greatest lower bound is also commonly used. The supremum of a subset of a partially ordered set is the least element in that is greater than or equal to each element of if such an element exists. In other words, it is the least element of that is greater than or equal to the greatest element of . Consequently, the supremum is also referred to as the least upper bound.

In mathematics, a (real) interval is the set of all real numbers lying between two fixed endpoints with no "gaps". Each endpoint is either a real number or positive or negative infinity, indicating the interval extends without a bound. An interval can contain neither endpoint, either endpoint, or both endpoints.

In the mathematical field of real analysis, the monotone convergence theorem is any of a number of related theorems proving the convergence of monotonic sequences that are also bounded. Informally, the theorems state that if a sequence is increasing and bounded above by a supremum, then the sequence will converge to the supremum; in the same way, if a sequence is decreasing and is bounded below by an infimum, it will converge to the infimum.

In mathematics, Hilbert's Nullstellensatz is a theorem that establishes a fundamental relationship between geometry and algebra. This relationship is the basis of algebraic geometry. It relates algebraic sets to ideals in polynomial rings over algebraically closed fields. This relationship was discovered by David Hilbert, who proved the Nullstellensatz in his second major paper on invariant theory in 1893.

In mathematics, Fatou's lemma establishes an inequality relating the Lebesgue integral of the limit inferior of a sequence of functions to the limit inferior of integrals of these functions. The lemma is named after Pierre Fatou.

<span class="mw-page-title-main">Maximal and minimal elements</span> Element that is not ≤ (or ≥) any other element

In mathematics, especially in order theory, a maximal element of a subset of some preordered set is an element of that is not smaller than any other element in . A minimal element of a subset of some preordered set is defined dually as an element of that is not greater than any other element in .

<span class="mw-page-title-main">Lindemann–Weierstrass theorem</span> On algebraic independence of exponentials of linearly independent algebraic numbers over Q

In transcendental number theory, the Lindemann–Weierstrass theorem is a result that is very useful in establishing the transcendence of numbers. It states the following:

In mathematics, specifically order theory, a well-quasi-ordering or wqo on a set is a quasi-ordering of for which every infinite sequence of elements from contains an increasing pair with

In mathematics, a subset of a preordered set is said to be cofinal or frequent in if for every it is possible to find an element in that is "larger than ".

In functional analysis and related areas of mathematics, a sequence space is a vector space whose elements are infinite sequences of real or complex numbers. Equivalently, it is a function space whose elements are functions from the natural numbers to the field K of real or complex numbers. The set of all such functions is naturally identified with the set of all possible infinite sequences with elements in K, and can be turned into a vector space under the operations of pointwise addition of functions and pointwise scalar multiplication. All sequence spaces are linear subspaces of this space. Sequence spaces are typically equipped with a norm, or at least the structure of a topological vector space.

In measure theory, an area of mathematics, Egorov's theorem establishes a condition for the uniform convergence of a pointwise convergent sequence of measurable functions. It is also named Severini–Egoroff theorem or Severini–Egorov theorem, after Carlo Severini, an Italian mathematician, and Dmitri Egorov, a Russian physicist and geometer, who published independent proofs respectively in 1910 and 1911.

In commutative algebra, an element b of a commutative ring B is said to be integral over a subring A of B if b is a root of some monic polynomial over A.

<span class="mw-page-title-main">Quasi-isometry</span> Function between two metric spaces that only respects their large-scale geometry

In mathematics, a quasi-isometry is a function between two metric spaces that respects large-scale geometry of these spaces and ignores their small-scale details. Two metric spaces are quasi-isometric if there exists a quasi-isometry between them. The property of being quasi-isometric behaves like an equivalence relation on the class of metric spaces.

Gordan's lemma is a lemma in convex geometry and algebraic geometry. It can be stated in several ways.

References

  1. 1 2 Dickson, L. E. (1913), "Finiteness of the odd perfect and primitive abundant numbers with n distinct prime factors", American Journal of Mathematics, 35 (4): 413–422, doi:10.2307/2370405, JSTOR   2370405 .
  2. 1 2 Buchberger, Bruno; Winkler, Franz (1998), Gröbner Bases and Applications, London Mathematical Society Lecture Note Series, vol. 251, Cambridge University Press, p. 83, ISBN   9780521632980 .
  3. Kruskal, Joseph B. (1972). "The theory of well-quasi-ordering: A frequently discovered concept". Journal of Combinatorial Theory . Series A. 13 (3): 298. doi: 10.1016/0097-3165(72)90063-5 .
  4. 1 2 Figueira, Diego; Figueira, Santiago; Schmitz, Sylvain; Schnoebelen, Philippe (2011), "Ackermannian and primitive-recursive bounds with Dickson's lemma", 26th Annual IEEE Symposium on Logic in Computer Science (LICS 2011), IEEE Computer Soc., Los Alamitos, CA, p. 269, arXiv: 1007.2989 , doi:10.1109/LICS.2011.39, MR   2858898, S2CID   9178090 .
  5. Onn, Shmuel (2008), "Convex Discrete Optimization", in Floudas, Christodoulos A.; Pardalos, Panos M. (eds.), Encyclopedia of Optimization, Vol. 1 (2nd ed.), Springer, pp. 513–550, arXiv: math/0703575 , Bibcode:2007math......3575O, ISBN   9780387747583 .
  6. Berlekamp, Elwyn R.; Conway, John H.; Guy, Richard K. (2003), "18 The Emperor and his Money", Winning Ways for your Mathematical Plays, Vol. 3 , Academic Press, pp. 609–640. See especially "Are outcomes computable", p. 630.