Fact-checked by Grok 2 weeks ago

Maximal and minimal elements

In , particularly within the field of , maximal and minimal elements are fundamental concepts in partially ordered sets (posets), where a partial is a binary that is reflexive, antisymmetric, and transitive. A maximal element in a poset is an element m such that no other element x in the set satisfies m < x, meaning m is not strictly less than any other element. Similarly, a minimal element is an element n such that no other element y satisfies y < n, indicating n has no element strictly below it in the . These elements play a crucial role in analyzing the structure of posets, as every finite nonempty poset contains at least one maximal and one minimal element, though there may be multiple such elements depending on the . Maximal and minimal elements must be distinguished from greatest (or maximum) and least (or minimum) elements: a greatest element is a maximal element that is greater than or equal to every element in the poset, while a least element is a minimal element less than or equal to all others. If a greatest element exists in a poset, it is unique and serves as the only maximal element, with an analogous uniqueness holding for the least element. In infinite posets, such as the set of integers under the usual order, there may be no maximal or minimal elements at all. These concepts extend beyond pure mathematics into applications like optimization problems, where maximal elements correspond to local maxima in partially ordered search spaces, and in combinatorics for studying lattices and chains. , which visualize posets by omitting transitive edges, are commonly used to identify maximal and minimal elements as those at the "top" and "bottom" layers without upward or downward connections, respectively.

Core Definitions

Partial orders and posets

A partial order on a set S is a binary relation \preceq that is reflexive, antisymmetric, and transitive. Reflexivity means that for every x \in S, x \preceq x; antisymmetry implies that if x \preceq y and y \preceq x, then x = y; and transitivity states that if x \preceq y and y \preceq z, then x \preceq z. A partially ordered set, or poset, consists of a set S together with a partial order \preceq on it, denoted (S, \preceq). Posets generalize total orders, allowing some elements to be incomparable, meaning neither x \preceq y nor y \preceq x holds for certain pairs. A common visual representation of a finite poset is the Hasse diagram, which depicts elements as vertices and draws edges only for covering relations—where x covers y if x > y and no z satisfies y < z < x—with transitive edges omitted and the order directed upward. Simple examples of posets include the power set \mathcal{P}(X) of a finite set X, ordered by inclusion \subseteq, where subsets are comparable if one is contained in the other. Another is the set of natural numbers \mathbb{N} under divisibility, where m \preceq n if m divides n. These structures form the foundational framework for , with all discussions of extrema in relying on this setup.

Maximal and minimal elements

In a (poset), maximal and minimal elements represent local extrema with respect to the order relation. Formally, let (P, \leq) be a poset. An element m \in P is maximal if there is no element in P strictly greater than m, that is, for all x \in P, if x \geq m then x = m. Equivalently, m is maximal if \neg \exists x \in P such that m < x. Dually, an element n \in P is minimal if there is no element in P strictly less than n, that is, for all x \in P, if x \leq n then x = n. Equivalently, n is minimal if \neg \exists x \in P such that x < n. These definitions highlight the local nature of maximal and minimal elements, as they require only that no strictly larger (or smaller) element exists, without demanding comparability to every other element in the poset. Consider the poset of all non-empty subsets of \{1, 2\} ordered by inclusion (\subseteq). The singletons \{1\} and \{2\} are minimal elements, since no proper non-empty subset is contained in them. In this poset, the full set \{1, 2\} is the unique maximal element. For a different example, take the poset of positive integers greater than 1 under divisibility (|), where a \leq b if a divides b. The prime numbers (such as 2, 3, 5) serve as minimal elements, as no integer greater than 1 strictly divides a prime. These examples illustrate how multiple maximal or minimal elements can coexist in a poset due to incomparabilities. Unlike greatest and least elements, which are comparable to every element in the poset, maximal and minimal elements need only lack successors or predecessors locally, allowing for several such elements in non-total orders.

Global Extrema

Greatest and least elements

In a partially ordered set (poset) (P, \leq), a greatest element g \in P is defined as an element such that x \leq g for every x \in P. Symmetrically, a least element \ell \in P satisfies \ell \leq x for every x \in P. These elements represent global maxima and minima, respectively, as they are comparable to all other elements in the poset via the order relation. If a greatest element exists in a poset, it is unique; this follows from the antisymmetry of the partial order, since supposing two greatest elements g and g' would imply g \leq g' and g' \leq g, hence g = g'. The same uniqueness holds for the least element. Unlike maximal or minimal elements, which may not be comparable to every other element, greatest and least elements are comparable to the entire poset. A classic example occurs in the power set \mathcal{P}(A) of a set A ordered by inclusion: the empty set \emptyset serves as the least element, since \emptyset \subseteq X for all X \subseteq A, while A itself is the greatest element, as X \subseteq A for all X \subseteq A. In the set of real numbers \mathbb{R} under the usual order \leq, no greatest or least element exists, but extending to the affinely extended real numbers \overline{\mathbb{R}} = \mathbb{R} \cup \{-\infty, +\infty\} introduces +\infty as the greatest element and -\infty as the least.

Existence and uniqueness

In a partially ordered set (poset), the greatest element, if it exists, is unique. To see this, suppose g and g' are both greatest elements. Then g \leq g' and g' \leq g by definition, so antisymmetry implies g = g'. The proof for the uniqueness of the least element is symmetric. The existence of greatest and least elements is not guaranteed in arbitrary posets. A finite counterexample is an antichain consisting of more than one element under the discrete order (where no two distinct elements are comparable); every element is both maximal and minimal, but there is no greatest or least element. An infinite counterexample is the poset (\mathbb{Q}, \leq) of rational numbers under the usual order, which has no greatest element because it is unbounded above (for any q \in \mathbb{Q}, there exists q' > q) and no least element because it is unbounded below (for any q \in \mathbb{Q}, there exists q'' < q). Sufficient conditions for existence include structural properties of the poset. Every nonempty finite lattice has a greatest element (the join of all its elements) and a least element (the meet of all its elements). More generally, every complete lattice has a greatest element, known as the , which is the supremum of the entire lattice, and a least element, known as the , which is the infimum of the entire lattice. In a complete lattice that is bounded above, the supremum of the poset serves as the greatest element; dually, if bounded below, the infimum serves as the least element.

Distinctions and Properties

Differences between maximal/minimal and greatest/least

In a partially ordered set (poset), the concepts of maximal and greatest elements, while related, differ fundamentally in their relational requirements. A greatest element g satisfies x \leq g for every element x in the poset, making it comparable to and above all others. In contrast, a maximal element m is one for which there exists no y such that m < y, meaning no strictly larger element exists above it, but m need not be comparable to or greater than every element in the set. This distinction arises because posets allow incomparability, unlike total orders where maximal and greatest elements coincide. The comparability issue highlights why multiple maximal elements can exist while a greatest element, if present, is unique. In a poset without total comparability, elements may form separate "branches" where each has no larger element above it but none dominates the others; thus, several maximals are possible, but no single greatest element exists unless one is above all. For instance, consider the poset consisting of \emptyset, \{1\}, and \{2\} under inclusion: the singletons \{1\} and \{2\} are both maximal (neither properly contains the other and nothing above them), but there is no greatest element. Hasse diagrams visually illustrate these differences: in a diagram with multiple disconnected upper branches, each tip represents a maximal element without a single greatest at the top, whereas a diagram converging to one apex shows a unique greatest element that is also maximal. For example, a Hasse diagram for the divisor poset of 6 (elements 1, 2, 3, 6) has 6 as the greatest (above all), but removing 6 leaves 2 and 3 as incomparable maximals. Every greatest element is maximal, but the converse does not hold: if g is greatest, then for any y \neq g, g \not< y since y \leq g, so no larger element exists; however, a maximal m may fail to be greatest if some incomparable z exists with z \not\leq m. This one-way implication underscores the broader applicability of maximal elements in non-linear structures.

Basic properties

In a finite partially ordered set (poset), maximal and minimal elements always exist. This follows from the finiteness of the poset: starting from any element, repeatedly moving to a strictly greater (or lesser) element would eventually terminate due to the absence of infinite chains, yielding a maximal (or minimal) element. A fundamental duality holds between maximal and minimal elements in posets. Specifically, the minimal elements of a poset (P, \leq) are precisely the maximal elements of the dual poset (P, \geq), obtained by reversing the order relation. This duality principle extends to all order-theoretic properties, ensuring that statements about maximals in one order correspond to statements about minimals in the reverse. Removing a maximal element from a poset produces a new poset in which the set of maximal elements may change, potentially introducing previously non-maximal elements as new maximals. For instance, in a poset with a unique maximal element, its removal yields a subposet whose maximal elements were covered by the original one. This property is key in inductive constructions and decompositions of posets. In posets, the existence of minimal elements in every nonempty subset is tied to well-foundedness. A poset is well-founded if every nonempty subset has a minimal element with respect to the order; the absence of infinite descending chains ensures this, as any descending sequence admits a least element. Conversely, non-well-founded posets may lack minimal elements in certain subsets due to infinite descents.

Advanced Structures

Directed sets

In order theory, a directed set is a partially ordered set (poset) in which every pair of elements has an upper bound within the set. Formally, given a poset (D, \leq), it is upward directed if for all a, b \in D, there exists c \in D such that a \leq c and b \leq c. This property extends to finite subsets, ensuring that the order structure allows for consistent "ascent" without isolated components. Downward directed sets are defined dually, where every pair has a lower bound. The absence of maximal elements plays a key role in the structure of infinite directed sets. In a directed set, any maximal element must coincide with the greatest element, because for any other element x, an upper bound y for both x and the maximal m satisfies y \geq m, but maximality of m implies y = m, hence m \geq x. Thus, directed sets lacking a greatest element have no maximal elements, enabling unbounded ascent: sequences of elements can continually increase without reaching a top. This contrasts with finite directed sets, which always possess a greatest (and hence maximal) element. A classic example is the set of natural numbers \mathbb{N} under the usual order \leq, which is upward directed since \max(n, m) serves as an upper bound for any n, m \in \mathbb{N}, yet it contains no maximal element due to its infinite, unbounded nature. Directed sets are essential in topology for defining , which generalize sequences: a net is a function from a directed set D to a topological space X, converging if neighborhoods shrink along the directed order, capturing limits in non-first-countable spaces. Directed sets connect to filters through their structural analogy in subset lattices. A prefilter, or filter base, is a downward-directed family of non-empty subsets under inclusion—meaning for any U, V in the family, there exists W with W \subseteq U \cap V—often constructed without minimal elements to generate proper filters that avoid the entire space. This downward direction mirrors the upward direction of general directed sets, positioning them as foundational prefilters in the poset of subsets when minimal elements are absent, facilitating applications in convergence and ultrafilter constructions.

Zorn's lemma

Zorn's lemma states that if every chain in a partially ordered set (P, \leq) has an upper bound in P, then P contains a maximal element. Here, a chain is a totally ordered subset, meaning any two elements are comparable under \leq, and an upper bound for a subset S \subseteq P is an element u \in P such that s \leq u for all s \in S. A standard proof sketch proceeds by applying the lemma recursively to the collection of all chains in P, partially ordered by inclusion. Every chain of chains (i.e., a collection of chains where any two are comparable by inclusion) has an upper bound given by their union, which remains a chain in P and lies in P. By the lemma's hypothesis (applied meta-theoretically), this collection of chains has a maximal element C, a maximal chain in P. Any element x \in C is then maximal in P, for if there existed y \in P with x < y, adjoining y to C would yield a larger chain, contradicting maximality of C. An alternative informal sketch assumes no maximal element exists and constructs an infinite strictly ascending chain x_1 < x_2 < \cdots by iteratively extending any finite chain (using the axiom of choice implicitly), yielding a chain without upper bound and contradicting the hypothesis. Zorn's lemma was introduced by Max Zorn in 1935 as a tool to streamline proofs relying on the axiom of choice or well-ordering principle, amid debates over non-constructive methods in set theory. In modern set theory, specifically Zermelo–Fraenkel set theory without the axiom of choice (ZF), Zorn's lemma is equivalent to the axiom of choice: each implies the other, and both are independent of ZF. Zorn's lemma underpins several existence results in algebra and beyond. For instance, it proves that every nonzero vector space over a field has a , by considering the poset of linearly independent subsets ordered by inclusion—every chain has an upper bound via union, so a maximal linearly independent set exists, spanning the space. Similarly, it establishes that every field admits an : the poset of algebraic extensions (or maximal subfields closed under roots) ordered by inclusion satisfies the chain condition, yielding a maximal element that is algebraically closed.

Applications and Examples

Consumer theory

In consumer theory, preferences over commodity bundles are modeled as a binary relation \succeq on the consumption set X \subseteq \mathbb{R}^n_+, typically assumed to be reflexive and transitive, forming a preorder (or partial order if antisymmetric). This relation captures the consumer's ranking of bundles, where x \succeq y indicates that bundle x is at least as preferred as y. The feasible choices are restricted to the budget set B(p, w) = \{x \in X : p \cdot x \leq w\}, where p \gg 0 denotes prices and w > 0 denotes . The consumer's demand consists of the maximal elements of B(p, w) under \succeq, namely those bundles x \in B(p, w) such that no y \in B(p, w) satisfies y \succ x (where \succ is the strict part of \succeq). These maximal elements represent the optimal choices, as they are undominated within the . Maximal elements in this context align with Pareto-efficient bundles from the individual consumer's viewpoint, meaning no alternative bundle in the set Pareto-dominates them by being strictly preferred without trade-offs in other dimensions (under the partial order induced by \succeq). If preferences are complete and transitive, the set of maximal elements is nonempty under standard and assumptions on B(p, w), ensuring the existence of demanded bundles. This framework underpins the approach, where observed choices rationalize a preference relation if they coincide with its maximal elements across budget sets. For instance, in a two-good with linear line and convex s, the maximal elements occur at the points of tangency between the budget line and the highest attainable indifference curve, corresponding to utility-maximizing allocations. Minimal elements under \succeq in the budget set are bundles x \in B(p, w) such that no other y \in B(p, w) satisfies y \prec x (where \prec is the strict part of the reverse preference), representing the least preferred options with no strictly inferior alternatives within the budget set. These are rarely emphasized in theory, as rational agents avoid them, but they can illustrate boundary cases like extreme corner solutions in non-monotonic preferences. The focus on maximal elements ties directly to maximization when \succeq admits a continuous representation u, where demanded bundles solve \max_{x \in B(p, w)} u(x).

Lattice theory

In lattice theory, a lattice is defined as a partially ordered set (poset) (L, \leq) in which every pair of elements a, b \in L has a least upper bound, called the join and denoted a \vee b, and a greatest lower bound, called the meet and denoted a \wedge b. This structure ensures that finite subsets also possess joins and meets, extending the binary operations associatively. In bounded lattices, which include a greatest element (top, denoted $1) and a least element (bottom, denoted &#36;0), these bounds serve as the unique maximal and minimal elements, respectively. Maximal elements in a are those with no strictly greater element, while minimal elements have no strictly smaller one; however, s often emphasize specialized forms like atoms and coatoms, particularly in modular s. An atom is a minimal element strictly above $0, meaning it covers &#36;0 (no element between them), and a coatom is dually a maximal element strictly below $1$. Join-irreducible elements, which cannot be expressed as the join of two strictly smaller elements, include atoms in atomic s and play a key role in decompositions; dually, meet-irreducible elements relate to coatoms in coatomic s. In modular s, such as vector subspaces under inclusion, atoms correspond to one-dimensional subspaces (maximal below certain joins), while coatoms are hyperplanes (maximal below the full space). A prominent example is the lattice of a $2^X for a X, ordered by inclusion, where the $0 is the [empty set](/page/Empty_set) and the top &#36;1 is X. Here, atoms are the singletons \{x\} for x \in X, serving as minimal non-bottom elements, while coatoms are the complements X \setminus \{x\}, which are maximal elements directly below $1.[25] Another example is the divisor [lattice](/page/Lattice) of a positive [integer](/page/Integer) n, consisting of divisors of n ordered by divisibility, with [bottom](/page/Bottom) &#36;1 and top n; maximal elements below n are the maximal proper divisors (e.g., for n=12, they include $6 and &#36;4), and minimal elements above $1 are the prime divisors like &#36;2 and $3$. In distributive lattices, where joins and meets satisfy a \wedge (b \vee c) = (a \wedge b) \vee (a \wedge c) and the dual, maximal ideals—principal ideals generated by maximal elements below $1—coincide with prime ideals, which are ideals Isuch that for anya, b \in L, if a \vee b \in Ithena \in Iorb \in I. This correspondence allows representation of the lattice via homomorphisms to the two-element lattice {0,1}$, with prime ideals as kernels, facilitating and duality in structures like Boolean algebras.

Bounds and ideals

In a partially ordered set (P, \leq), an upper bound of a subset S \subseteq P is an b \in P such that s \leq b for every s \in S. Dually, a lower bound of S is an l \in P such that l \leq s for every s \in S. Ideals in a poset provide a structure closely related to lower bounds, as they are down-sets: subsets I \subseteq P that are closed downward, meaning if x \in I and y \leq x with y \in P, then y \in I. A is generated by a single a \in P and consists of all lower bounds of a, formally \downarrow a = \{ x \in P \mid x \leq a \}. When a is a minimal element of P, the principal ideal \downarrow a is the singleton \{a\}, as no distinct is strictly less than a. In the poset of all ideals of P, ordered by inclusion, a maximal ideal is an ideal I that is proper (i.e., I \neq P) and not properly contained in any larger ideal of P; this notion is dual to that of maximal elements in the original poset P. For example, in ring theory, the ideals of a commutative ring R form a poset under inclusion, and a maximal ideal \mathfrak{m} of R is one such that no proper ideal of R strictly contains \mathfrak{m} except R itself, with the quotient ring R / \mathfrak{m} being a field.

Chains and antichains

In a partially ordered set (poset), a is a in which every pair of distinct s is comparable under the partial order, meaning the subset is . A maximal chain is a chain that cannot be properly extended by adding another element from the poset while preserving the total order property; thus, no superset of a maximal chain is itself a chain. An , in contrast, is a of the poset consisting of pairwise elements, where no two distinct elements are related by the partial order. A maximal antichain is one that is not properly contained in any larger antichain; it represents a "widest" level of incomparability in the poset, though it may not be the largest possible antichain. The size of the largest antichain in a finite poset is known as the width of the poset. Dilworth's theorem provides a fundamental connection between chains and antichains in finite posets: the width of the poset (the of a maximum-sized ) equals the minimum number of needed to the entire poset into disjoint . This result, originally proved using a decomposition argument, implies that any finite poset can be covered by a chain whose size matches its maximum , with applications in and optimization. Chains and antichains relate directly to maximal and minimal elements. In any , since all are pairwise incomparable, each is both a maximal element and a minimal element within the subset itself. For a maximal in a finite poset, the least element of the is a minimal element of the poset (as no smaller element can be added below it), and the greatest is a maximal element of the poset (as no larger element can be added above it).

References

  1. [1]
    [PDF] 4. Partial Orderings - FSU Mathematics
    Maximal and Minimal Elements. Definition 4.11.1. Let (A, R) be a poset. (1) An element a ∈ A is a minimal element if there does not exist an element b ∈ A ...
  2. [2]
    [PDF] Math 3012 – Applied Combinatorics Lecture 13 - William T. Trotter
    Oct 1, 2015 · Definition An element x of a poset P is said to be a maximal point of P when there is no point y of P with y > x in P. Definition An element w ...
  3. [3]
    2.2: Equivalence Relations, and Partial order - Mathematics LibreTexts
    Nov 21, 2024 · Definition: Partial Order. A binary relation is a partial order if and only if the relation is reflexive(R), antisymmetric(A) and transitive(T) ...
  4. [4]
    4. Partial Orders - Random Services
    A partial order on a set \(S\) is a relation \(\preceq\) on \(S\) that is reflexive, anti-symmetric, and transitive. The pair \( (S, \preceq) \) is called a ...
  5. [5]
    [PDF] Lecture 7 1 Partially ordered sets
    Feb 24, 2011 · Definition 5 The Hasse diagram of a partially ordered set P is the (directed) graph whose vertices are the elements of P and whose edges are the ...
  6. [6]
    Partial Orders and Lattices (Set-2) | Mathematics - GeeksforGeeks
    Jul 11, 2025 · Examples of Partial Order Sets. Set of Natural Numbers with Divisibility: Consider the set of natural numbers {1,2,3,4,…}with the relation a ...
  7. [7]
    [PDF] partially ordered sets and lattices
    If (X,≤) is a finite partially ordered set, then X has a maximal and a minimal element. An element x ∈ X is maximal if x ≤ y implies x = y. Note there can be.
  8. [8]
    [PDF] Section 7.6 Partial Orderings Definition - Temple CIS
    Maximal and Minimal Elements. Definition: Let (A, R) be a poset. Then a in A is a minimal element if there does not exist an element b in A such that bRa ...
  9. [9]
    [PDF] PARTIALLY ORDERED SETS
    A partially ordered set or poset is a set P and a binary relation such that ... Choose a maximal element of P and label it pn. Assume that. (1) can be ...
  10. [10]
  11. [11]
    Affinely Extended Real Numbers -- from Wolfram MathWorld
    The set R union {+infty,-infty} obtained by adjoining two improper elements to the set R of real numbers is normally called the set of (affinely) extended real ...
  12. [12]
    [PDF] Preliminary Notes on Lattices 1 Partially ordered sets - P.J. Healy
    P be a partially ordered set of parameters. Let φ: P ↠ X be a constraint ... The greatest element is the unique maximal element of A. Proof: Suppose x ...
  13. [13]
    [PDF] Math 155 (Lecture 19)
    Oct 18, 2011 · We say that a ∈ A is a maximal element if b ≥ a implies b = a. Let A be a partially ordered set. If A has a least element a, then a is unique, ...
  14. [14]
    [PDF] Section 3.2 Lattices Let (A, ≤) be a (reflexive) order, considered ...
    Similarly, ∅ , if it exists, is the top element, or greatest element of the order; it is ... A complete lattice is an order (A, ≤) in which X , X exist.
  15. [15]
  16. [16]
    Maximal Element -- from Wolfram MathWorld
    Let (A,<=) be a partially ordered set. Then an element m in A is said to be maximal if, for all a in A, m!<=a. Alternatively, an element m in A is maximal ...
  17. [17]
    19.5: Maximal/minimal Elements - Mathematics LibreTexts
    Feb 19, 2022 · If a subset of a partially ordered set contains a maximum element, then that maximum element is unique. And similarly for a minimum element.Fact 19 . 5 . 1 · Example 19 . 5 . 2 : Maximums... · Test 19 . 5 . 1 : Maximal...
  18. [18]
    None
    Below is a merged summary of maximal and minimal elements in posets from Birkhoff's *Lattice Theory* (1948, Revised Edition), consolidating all information from the provided segments into a comprehensive response. To maximize density and clarity, I will use a table in CSV format to organize the details, followed by a narrative summary that ties together the key points not fully captured in the table. The table will cover definitions, existence, duality, properties (including removal and countability), and relevant sections/URLs, drawing from all segments.
  19. [19]
    Hook length property of d-complete posets via q-integrals
    If P has a unique maximal element, we define P − to be the poset obtained from P by removing the maximal element. Note that ( P + ) − = P for any poset P.
  20. [20]
    Chapter III. The Well-Founded Sets - ScienceDirect.com
    0 Thus, R is well-founded on A iff every non-empty subset has an Rminimal element. ... minimal element. For example, the empty relation, 0, is well-founded on any ...Missing: posets | Show results with:posets
  21. [21]
    directed set - PlanetMath
    Mar 22, 2013 · A directed set is a partially ordered set. (A,≤) such that whenever a,b∈A a , b ∈ A there is an x∈A x ∈ A such that a≤x a ≤ x and b≤x b ≤ x .Missing: theory | Show results with:theory<|control11|><|separator|>
  22. [22]
    Directed set - Encyclopedia of Mathematics
    Oct 14, 2014 · A set A with partial order ≤ is called upwards (respectively, downwards) directed if ≤ (respectively, the opposite order ≥) is a directed order.
  23. [23]
    [PDF] Applications of computability theory to partial and linear orders
    Note that in directed posets a maximal element is also a greatest element. ... Recall that the Tukey type of a directed set is 1 if that set has a greatest ...
  24. [24]
    [PDF] Zorn's lemma and some applications - Keith Conrad
    Zorn's lemma is not intuitive, but it is logically equivalent to more intuitively plausible statements in set theory like the Axiom of Choice, which says every ...
  25. [25]
  26. [26]
    [PDF] Section 0.7. The Axiom of Choice, Order, and Zorn's Lemma
    Jan 24, 2021 · Every vector space has a basis. • Every field F has an algebraic closure F. Note. The fact that every infinite set has a countable subset ...
  27. [27]
    Revealed Preference Theory - jstor
    It remains to show that h (B) is always the set of G-maximal elements of B. We first show that h(B)c {z: zeB & VvVGBzGv}. Let xe-h(B). Then, clearly ...
  28. [28]
    [PDF] Notes on Lattice Theory J. B. Nation University of Hawaii
    An element q of a lattice L is called join irreducible if q = "F for a finite set F implies q ∈ F, i.e., q is not the join of other elements. The set of all ...
  29. [29]
    [PDF] BIRKHOFF 1948 Lattice Theory Revised Edition - Chapman University
    Lattice theory, developed in three stages, has applications to algebra, geometry, set theory, and functional analysis. It is a substantial branch of modern ...
  30. [30]
    [PDF] The Maximal Ideal Theorem for Lattices of Sets
    It is a well-known fact that a maximal ideal of a distributive lattice must be prime; the converse is easily shown to be false in general. Now it is easy to ...
  31. [31]
    [PDF] Lattice Theory Lecture 2 Distributive lattices - nmsu math
    Prime ideals. Definition Let 2 be the 2-element lattice. Proposition For P a prime ideal of a distributive lattice D, there is a homomorphism ϕp ∶ D → 2 ...
  32. [32]
    upper bound - PlanetMath.org
    Mar 22, 2013 · Let S be a set with a partial ordering ≤ , and let T be a subset of S . An upper bound. for T is an element z∈S z ∈ S such that x≤z x ≤ z for ...
  33. [33]
    order ideal - PlanetMath
    Mar 22, 2013 · An order ideal is also called an ideal for short. An ideal is said to be principal if it has the form ↓x ↓ x for some x∈P x ∈ P.Missing: theory | Show results with:theory
  34. [34]
    maximal ideal - PlanetMath
    Mar 22, 2013 · , an ideal m⊂R 𝔪 ⊂ R is maximal if and only if the quotient ring R/m R / 𝔪 is a field. Title, maximal ideal.
  35. [35]
    Maximal Ideal -- from Wolfram MathWorld
    A maximal ideal of a ring R is an ideal I, not equal to R, such that there are no ideals in between I and R.
  36. [36]
    Partial Order -- from Wolfram MathWorld
    For a partial order, the size of the longest chain (antichain) is called the partial order length (partial order width). A partially ordered set is also called ...
  37. [37]
    [PDF] Math 3012 – Applied Combinatorics Lecture 14 - William T. Trotter
    Oct 6, 2015 · Definition A chain is maximal when no superset is also a chain. Page 8. Height of a Poset. Definition The height of a poset P is the.
  38. [38]
    Antichain -- from Wolfram MathWorld
    Let P be a finite partially ordered set, then an antichain in P is a set of pairwise incomparable elements. Antichains are also called Sperner systems in ...<|separator|>
  39. [39]
    [PDF] A Decomposition Theorem for Partially Ordered Sets - UCSD Math
    A Decomposition Theorem for Partially Ordered Sets. Author(s): R. P. Dilworth. Source: Annals of Mathematics, Second Series, Vol. 51, No. 1 (Jan., 1950), pp ...
  40. [40]
    A Decomposition Theorem for Partially Ordered Sets - jstor
    A DECOMPOSITION THEOREM FOR PARTIALLY ORDERED SETS. BY R. P. DILWORTH. (Received August 23, 1948). 1. Introduction. Let P be a partially ordered set. Two ...