Modular lattice

Last updated
A modular lattice of order dimension 2. As with all finite 2-dimensional lattices, its Hasse diagram is an st-planar graph. 2d modular lattice.svg
A modular lattice of order dimension 2. As with all finite 2-dimensional lattices, its Hasse diagram is an st-planar graph.

In the branch of mathematics called order theory, a modular lattice is a lattice that satisfies the following self-dual condition,

Contents

Modular law
ab implies a ∨ (xb) = (ax) ∧ b

where x, a, b are arbitrary elements in the lattice,    is the partial order, and    and   (called join and meet respectively) are the operations of the lattice. This phrasing emphasizes an interpretation in terms of projection onto the sublattice [a, b], a fact known as the diamond isomorphism theorem. [1] An alternative but equivalent condition stated as an equation (see below) emphasizes that modular lattices form a variety in the sense of universal algebra.

Modular lattices arise naturally in algebra and in many other areas of mathematics. In these scenarios, modularity is an abstraction of the 2nd Isomorphism Theorem. For example, the subspaces of a vector space (and more generally the submodules of a module over a ring) form a modular lattice.

In a not necessarily modular lattice, there may still be elements b for which the modular law holds in connection with arbitrary elements x and a (for ab). Such an element is called a right modular element. Even more generally, the modular law may hold for any a and a fixed pair (x, b). Such a pair is called a modular pair, and there are various generalizations of modularity related to this notion and to semimodularity.

Modular lattices are sometimes called Dedekind lattices after Richard Dedekind, who discovered the modular identity in several motivating examples.

Introduction

The modular law can be seen as a restricted associative law that connects the two lattice operations similarly to the way in which the associative law λ(μx) = (λμ)x for vector spaces connects multiplication in the field and scalar multiplication.

The restriction ab is clearly necessary, since it follows from a ∨ (xb) = (ax) ∧ b. In other words, no lattice with more than one element satisfies the unrestricted consequent of the modular law.

It is easy to see [2] that ab implies a ∨ (xb) ≤ (ax) ∧ b in every lattice. Therefore, the modular law can also be stated as

Modular law (variant)
ab implies (ax) ∧ ba ∨ (xb).

The modular law can be expressed as an equation that is required to hold unconditionally. Since ab implies a = ab and since abb, replace a with ab in the defining equation of the modular law to obtain:

Modular identity
(ab) ∨ (xb) = ((ab) ∨ x) ∧ b.

This shows that, using terminology from universal algebra, the modular lattices form a subvariety of the variety of lattices. Therefore, all homomorphic images, sublattices and direct products of modular lattices are again modular.

Examples

N5, the smallest non-modular lattice: x[?](a[?]b) = x[?]0 = x [?] b = 1[?]b =(x[?]a)[?]b. Smallest nonmodular lattice 2.svg
N5, the smallest non-modular lattice: x∨(ab) = x∨0 = xb = 1∧b =(xa)∧b.

The lattice of submodules of a module over a ring is modular. As a special case, the lattice of subgroups of an abelian group is modular.

The lattice of normal subgroups of a group is modular. But in general the lattice of all subgroups of a group is not modular. For an example, the lattice of subgroups of the dihedral group of order 8 is not modular.

The smallest non-modular lattice is the "pentagon" lattice N5 consisting of five elements 0, 1, x, a, b such that 0 < x < b < 1, 0 < a < 1, and a is not comparable to x or to b. For this lattice,

x ∨ (ab) = x ∨ 0 = x < b = 1 ∧ b = (xa) ∧ b

holds, contradicting the modular law. Every non-modular lattice contains a copy of N5 as a sublattice. [3]

Properties

Every distributive lattice is modular. [4] [5]

Dilworth (1954) proved that, in every finite modular lattice, the number of join-irreducible elements equals the number of meet-irreducible elements. More generally, for every k, the number of elements of the lattice that cover exactly k other elements equals the number that are covered by exactly k other elements. [6]

A useful property to show that a lattice is not modular is as follows:

A lattice G is modular if and only if, for any a, b, cG,

Sketch of proof: Let G be modular, and let the premise of the implication hold. Then using absorption and modular identity:

c = (cb) ∨ c = (ab) ∨ c = a ∧ (bc) = a ∧ (ba) = a

For the other direction, let the implication of the theorem hold in G. Let a,b,c be any elements in G, such that ca. Let x = (ab) ∨ c, y = a ∧ (bc). From the modular inequality immediately follows that xy. If we show that xb = yb, xb = yb, then using the assumption x = y must hold. The rest of the proof is routine manipulation with infima, suprema and inequalities.[ citation needed ]

Diamond isomorphism theorem

For any two elements a,b of a modular lattice, one can consider the intervals [ab, b] and [a, ab]. They are connected by order-preserving maps

φ: [ab, b] → [a, ab] and
ψ: [a, ab] → [ab, b]

that are defined by φ(x) = xa and ψ(y) = yb.

The composition ψφ is an order-preserving map from the interval [ab, b] to itself which also satisfies the inequality ψ(φ(x)) = (xa) ∧ bx. The example shows that this inequality can be strict in general. In a modular lattice, however, equality holds. Since the dual of a modular lattice is again modular, φψ is also the identity on [a, ab], and therefore the two maps φ and ψ are isomorphisms between these two intervals. This result is sometimes called the diamond isomorphism theorem for modular lattices. A lattice is modular if and only if the diamond isomorphism theorem holds for every pair of elements.

The diamond isomorphism theorem for modular lattices is analogous to the second isomorphism theorem in algebra, and it is a generalization of the lattice theorem.

The centred hexagon lattice S7, also known as D2, is M-symmetric but not modular. Centred hexagon lattice D2.svg
The centred hexagon lattice S7, also known as D2, is M-symmetric but not modular.

In any lattice, a modular pair is a pair (a, b) of elements such that for all x satisfying a  bx  b, we have (x  a)  b = x, i.e. if one half of the diamond isomorphism theorem holds for the pair. [7] An element b of a lattice is called a right modular element if (a, b) is a modular pair for all elements a, and an element a is called a left modular element if (a, b) is a modular pair for all elements b. [8]

A lattice with the property that if (a, b) is a modular pair, then (b, a) is also a modular pair is called an M-symmetric lattice. [9] Thus, in an M-symmetric lattice, every right modular element is also left modular, and vice-versa. Since a lattice is modular if and only if all pairs of elements are modular, clearly every modular lattice is M-symmetric. In the lattice N5 described above, the pair (b, a) is modular, but the pair (a, b) is not. Therefore, N5 is not M-symmetric. The centred hexagon lattice S7 is M-symmetric but not modular. Since N5 is a sublattice of S7, it follows that the M-symmetric lattices do not form a subvariety of the variety of lattices.

M-symmetry is not a self-dual notion. A dual modular pair is a pair which is modular in the dual lattice, and a lattice is called dually M-symmetric or M*-symmetric if its dual is M-symmetric. It can be shown that a finite lattice is modular if and only if it is M-symmetric and M*-symmetric. The same equivalence holds for infinite lattices which satisfy the ascending chain condition (or the descending chain condition).

Several less important notions are also closely related. A lattice is cross-symmetric if for every modular pair (a, b) the pair (b, a) is dually modular. Cross-symmetry implies M-symmetry but not M*-symmetry. Therefore, cross-symmetry is not equivalent to dual cross-symmetry. A lattice with a least element 0 is ⊥-symmetric if for every modular pair (a, b) satisfying a  b = 0 the pair (b, a) is also modular.

History

Free modular lattice generated by three elements {x,y,z} Free modular lattice with 3 generators (x,y,z).gif
Free modular lattice generated by three elements {x,y,z}

The definition of modularity is due to Richard Dedekind, who published most of the relevant papers after his retirement. In a paper published in 1894[ citation needed ] he studied lattices, which he called dual groups (German : Dualgruppen) as part of his "algebra of modules" and observed that ideals satisfy what we now call the modular law. He also observed that for lattices in general, the modular law is equivalent to its dual.

In another paper in 1897, Dedekind studied the lattice of divisors with gcd and lcm as operations, so that the lattice order is given by divisibility. [10] In a digression he introduced and studied lattices formally in a general context. [10] :10–18 He observed that the lattice of submodules of a module satisfies the modular identity. He called such lattices dual groups of module type (Dualgruppen vom Modultypus). He also proved that the modular identity and its dual are equivalent. [10] :13

In the same paper, Dedekind also investigated the following stronger form [10] :14 of the modular identity, which is also self-dual: [10] :9

(xb) ∨ (ab) = [xa] ∧ b.

He called lattices that satisfy this identity dual groups of ideal type (Dualgruppen vom Idealtypus). [10] :13 In modern literature, they are more commonly referred to as distributive lattices. He gave examples of a lattice that is not modular and of a modular lattice that is not of ideal type. [10] :14

A paper published by Dedekind in 1900 had lattices as its central topic: He described the free modular lattice generated by three elements, a lattice with 28 elements (see picture). [11]

See also

Notes

  1. "Why are modular lattices important?". Mathematics Stack Exchange. Retrieved 2018-09-17.
  2. The following is true for any lattice: a ∨ (xb) ≤ (ax) ∧ (ab). Also, whenever ab, then ab = b.
  3. Blyth, T. S. (2005). "Modular lattices". Lattices and Ordered Algebraic Structures. Universitext. London: Springer. Theorem 4.4. doi:10.1007/1-84628-127-X_4. ISBN   978-1-85233-905-0.
  4. Blyth, T. S. (2005). "Modular lattices". Lattices and Ordered Algebraic Structures. Universitext. London: Springer. p. 65. doi:10.1007/1-84628-127-X_4. ISBN   978-1-85233-905-0.
  5. In a distributive lattice, the following holds: . Moreover, the absorption law, , is true for any lattice. Substituting this for the second conjunct of the right-hand side of the former equation yields the Modular Identity.
  6. Dilworth, R. P. (1954), "Proof of a conjecture on finite modular lattices", Annals of Mathematics , Second Series, 60 (2): 359–364, doi:10.2307/1969639, JSTOR   1969639, MR   0063348 . Reprinted in Bogart, Kenneth P.; Freese, Ralph; Kung, Joseph P. S., eds. (1990), "Proof of a Conjecture on Finite Modular Lattices", The Dilworth Theorems: Selected Papers of Robert P. Dilworth, Contemporary Mathematicians, Boston: Birkhäuser, pp. 219–224, doi:10.1007/978-1-4899-3558-8_21, ISBN   978-1-4899-3560-1
  7. The French term for modular pair is couple modulaire. A pair (a, b) is called a paire modulaire in French if both (a, b) and (b, a) are modular pairs.
  8. Modular element has been varying defined by different authors to mean right modular (Stern (1999 , p. 74)), left modular (Orlik & Terao (1992 , Definition 2.25)), both left and right modular (or dual right modular) (Sagan (1999), Schmidt (1994 , p. 43)), or satisfying a modular rank condition (Stanley (2004 , Definition 4.12)). These notions are equivalent in a semimodular lattice, but not in general.
  9. Some authors, e.g. Fofanova (2001), refer to such lattices as semimodular lattices. Since every M-symmetric lattice is semimodular and the converse holds for lattices of finite length, this can only lead to confusion for infinite lattices.
  10. 1 2 3 4 5 6 7 Dedekind, Richard (1897), "Über Zerlegungen von Zahlen durch ihre grössten gemeinsamen Theiler" (PDF), Festschrift der Herzogl. Technischen Hochschule Carolo-Wilhelmina bei Gelegenheit der 69. Versammlung Deutscher Naturforscher und Ärzte in Braunschweig, Friedrich Vieweg und Sohn
  11. Dedekind, Richard (1900), "Über die von drei Moduln erzeugte Dualgruppe", Mathematische Annalen, 53 (3): 371–403, doi:10.1007/BF01448979, S2CID   122529830

Related Research Articles

<span class="mw-page-title-main">Boolean algebra (structure)</span> Algebraic structure modeling logical operations

In abstract algebra, a Boolean algebra or Boolean lattice is a complemented distributive lattice. This type of algebraic structure captures essential properties of both set operations and logic operations. A Boolean algebra can be seen as a generalization of a power set algebra or a field of sets, or its elements can be viewed as generalized truth values. It is also a special case of a De Morgan algebra and a Kleene algebra.

In mathematics, a Heyting algebra (also known as pseudo-Boolean algebra) is a bounded lattice (with join and meet operations written ∨ and ∧ and with least element 0 and greatest element 1) equipped with a binary operation ab of implication such that (ca) ≤ b is equivalent to c ≤ (ab). From a logical standpoint, AB is by this definition the weakest proposition for which modus ponens, the inference rule AB, AB, is sound. Like Boolean algebras, Heyting algebras form a variety axiomatizable with finitely many equations. Heyting algebras were introduced by Arend Heyting (1930) to formalize intuitionistic logic.

In mathematics, a distributive lattice is a lattice in which the operations of join and meet distribute over each other. The prototypical examples of such structures are collections of sets for which the lattice operations can be given by set union and intersection. Indeed, these lattices of sets describe the scenery completely: every distributive lattice is—up to isomorphism—given as such a lattice of sets.

In mathematics, a quotient algebra is the result of partitioning the elements of an algebraic structure using a congruence relation. Quotient algebras are also called factor algebras. Here, the congruence relation must be an equivalence relation that is additionally compatible with all the operations of the algebra, in the formal sense described below. Its equivalence classes partition the elements of the given algebraic structure. The quotient algebra has these classes as its elements, and the compatibility conditions are used to give the classes an algebraic structure.

In mathematics, there is an ample supply of categorical dualities between certain categories of topological spaces and categories of partially ordered sets. Today, these dualities are usually collected under the label Stone duality, since they form a natural generalization of Stone's representation theorem for Boolean algebras. These concepts are named in honor of Marshall Stone. Stone-type dualities also provide the foundation for pointless topology and are exploited in theoretical computer science for the study of formal semantics.

This is a glossary of some terms used in various branches of mathematics that are related to the fields of order, lattice, and domain theory. Note that there is a structured list of order topics available as well. Other helpful resources might be the following overview articles:

A lattice is an abstract structure studied in the mathematical subdisciplines of order theory and abstract algebra. It consists of a partially ordered set in which every pair of elements has a unique supremum and a unique infimum. An example is given by the power set of a set, partially ordered by inclusion, for which the supremum is the union and the infimum is the intersection. Another example is given by the natural numbers, partially ordered by divisibility, for which the supremum is the least common multiple and the infimum is the greatest common divisor.

In mathematical order theory, an ideal is a special subset of a partially ordered set (poset). Although this term historically was derived from the notion of a ring ideal of abstract algebra, it has subsequently been generalized to a different notion. Ideals are of great importance for many constructions in order and lattice theory.

In the mathematical area of order theory, there are various notions of the common concept of distributivity, applied to the formation of suprema and infima. Most of these apply to partially ordered sets that are at least lattices, but the concept can in fact reasonably be generalized to semilattices as well.

<span class="mw-page-title-main">Complemented lattice</span>

In the mathematical discipline of order theory, a complemented lattice is a bounded lattice, in which every element a has a complement, i.e. an element b satisfying a ∨ b = 1 and a ∧ b = 0. Complements need not be unique.

<span class="mw-page-title-main">Lattice of subgroups</span>

In mathematics, the lattice of subgroups of a group is the lattice whose elements are the subgroups of , with the partial ordering being set inclusion. In this lattice, the join of two subgroups is the subgroup generated by their union, and the meet of two subgroups is their intersection.

In mathematics, a Riesz space, lattice-ordered vector space or vector lattice is a partially ordered vector space where the order structure is a lattice.

In abstract algebra, a skew lattice is an algebraic structure that is a non-commutative generalization of a lattice. While the term skew lattice can be used to refer to any non-commutative generalization of a lattice, since 1989 it has been used primarily as follows.

<span class="mw-page-title-main">Semimodular lattice</span>

In the branch of mathematics known as order theory, a semimodular lattice, is a lattice that satisfies the following condition:

<span class="mw-page-title-main">Dedekind–MacNeille completion</span> Smallest complete lattice containing a partial order

In mathematics, specifically order theory, the Dedekind–MacNeille completion of a partially ordered set is the smallest complete lattice that contains it. It is named after Holbrook Mann MacNeille whose 1937 paper first defined and constructed it, and after Richard Dedekind because its construction generalizes the Dedekind cuts used by Dedekind to construct the real numbers from the rational numbers. It is also called the completion by cuts or normal completion.

In the mathematics of matroids and lattices, a geometric lattice is a finite atomistic semimodular lattice, and a matroid lattice is an atomistic semimodular lattice without the assumption of finiteness. Geometric lattices and matroid lattices, respectively, form the lattices of flats of finite, or finite and infinite, matroids, and every geometric or matroid lattice comes from a matroid in this way.

In mathematics, Birkhoff's representation theorem for distributive lattices states that the elements of any finite distributive lattice can be represented as finite sets, in such a way that the lattice operations correspond to unions and intersections of sets. Here, a lattice is an abstract structure with two binary operations, the "meet" and "join" operations, which must obey certain axioms; it is distributive if these two operations obey the distributive law. The union and intersection operations, in a family of sets that is closed under these operations, automatically form a distributive lattice, and Birkhoff's representation theorem states that every finite distributive lattice can be formed in this way. It is named after Garrett Birkhoff, who published a proof of it in 1937.

In universal algebra and lattice theory, a tolerance relation on an algebraic structure is a reflexive symmetric relation that is compatible with all operations of the structure. Thus a tolerance is like a congruence, except that the assumption of transitivity is dropped. On a set, an algebraic structure with empty family of operations, tolerance relations are simply reflexive symmetric relations. A set that possesses a tolerance relation can be described as a tolerance space. Tolerance relations provide a convenient general tool for studying indiscernibility/indistinguishability phenomena. The importance of those for mathematics had been first recognized by Poincaré.

In mathematics and mathematical logic, Boolean algebra is a branch of algebra. It differs from elementary algebra in two ways. First, the values of the variables are the truth values true and false, usually denoted 1 and 0, whereas in elementary algebra the values of the variables are numbers. Second, Boolean algebra uses logical operators such as conjunction (and) denoted as , disjunction (or) denoted as , and the negation (not) denoted as ¬. Elementary algebra, on the other hand, uses arithmetic operators such as addition, multiplication, subtraction, and division. Boolean algebra is therefore a formal way of describing logical operations, in the same way that elementary algebra describes numerical operations.

In mathematics, a supersolvable lattice is a graded lattice that has a maximal chain of elements, each of which obeys a certain modularity relationship. The definition encapsulates many of the nice properties of lattices of subgroups of supersolvable groups.

References