Fact-checked by Grok 2 weeks ago

Well-ordering principle

The well-ordering principle states that every non-empty subset of the positive integers contains a least element. This property, often regarded as an for the natural numbers, ensures that the set of positive integers \mathbb{N}^+ = \{1, 2, 3, \dots\} (or nonnegative integers \mathbb{N} \cup \{0\}) is well-ordered under the standard ordering, meaning no infinite descending chains exist and every subset is bounded below by a minimal member. Equivalent to the principle of , the well-ordering principle provides a foundational tool for proving statements about the natural numbers by : assuming a non-empty set of counterexamples leads to the existence of a smallest counterexample, which can then be shown to satisfy the property, yielding a . This equivalence arises because establishes properties for all natural numbers starting from a base case, while well-ordering guarantees the minimal element needed to initiate such recursive arguments. In practice, it underpins proofs by minimal counterexample, such as demonstrating that every positive greater than 1 factors uniquely into primes or that the of the first n positive integers is \frac{n(n+1)}{2}. Distinct from the well-ordering theorem—which asserts that every set can be well-ordered and is equivalent to the in Zermelo-Fraenkel —the well-ordering principle specifically applies to the natural numbers and holds without requiring additional axioms beyond the . It plays a central role in and , enabling rigorous analysis of divisibility, Diophantine equations, and algorithmic termination, while highlighting the ordered structure that distinguishes the naturals from denser sets like or reals, which admit non-empty subsets without least elements (e.g., the positive rationals).

Definition and Fundamentals

Formal Statement

The well-ordering principle asserts that every non-empty subset of the natural numbers possesses a least element with respect to the standard ordering. Here, the natural numbers \mathbb{N} are taken to be the set of non-negative integers \{0, 1, 2, \dots\}. This principle is a foundational property of the natural numbers in and arithmetic. Formally, the principle can be stated as follows: For every non-empty S \subseteq \mathbb{N}, there exists an element m \in S such that m \leq n for all n \in S. A well-ordering on a set extends beyond a simple by requiring not only that every pair of distinct elements is comparable (i.e., for any a, b \in \mathbb{N} with a \neq b, either a \leq b or b \leq a), but also that every non-empty has a minimal element. This additional condition prevents infinite descending chains, such as would occur in the integers under the usual order. As an immediate consequence, the natural numbers \mathbb{N} themselves constitute a well-ordered set under the standard relation < (or equivalently \leq).

Historical Context

The well-ordering principle has roots traceable to ancient mathematics, where it was implicitly employed in proofs concerning natural numbers. Around 300 BCE, Euclid's Elements utilized arguments equivalent to the principle in demonstrations such as Proposition 31 of Book VII, which establishes properties of the greatest common divisor through a process akin to infinite descent, assuming no infinite decreasing sequence of positive integers exists. In the early 19th century, mathematicians like Bernhard Bolzano contributed to the rigorous foundations of analysis and arithmetic, laying groundwork for formalizing inductive reasoning and ordered structures, though explicit statements of well-ordering emerged later. By the late 19th century, Richard Dedekind provided a pivotal formalization in his 1888 essay Was sind und was sollen die Zahlen?, where Theorem 96 explicitly states that every non-empty subset of the natural numbers possesses a least element, serving as the basis for mathematical induction within his construction of the naturals via Dedekind cuts and chains. This work marked a shift toward axiomatic treatments of arithmetic, emphasizing the principle's role in proving fundamental properties without reliance on intuition. The principle's scope expanded dramatically in set theory through Ernst Zermelo's 1904 paper Beweis, daß jede Menge wohlgeordnet werden kann, which proved that every set can be well-ordered, relying on what became known as the axiom of choice. Building on Georg Cantor's earlier development of transfinite ordinals and well-ordered sets in the 1890s, this result elevated the well-ordering principle from an arithmetic intuition to a cornerstone of axiomatic set theory. Post-Cantor, it achieved axiomatic status in systems like Zermelo-Fraenkel set theory, where its equivalence to the axiom of choice underscored its foundational importance in modern mathematics.

Equivalences to Other Principles

Equivalence to Mathematical Induction

The principle of mathematical induction states that for any property P defined on the nonnegative integers, if P(0) holds and for every nonnegative integer n, P(n) implies P(n+1), then P(n) holds for all nonnegative integers n. This can be expressed formally as: P(0) \land \forall n \, (P(n) \to P(n+1)) \to \forall n \, P(n). The well-ordering principle, which asserts that every nonempty subset of the nonnegative integers has a least element, is logically equivalent to the principle of mathematical induction. To establish this equivalence, it must be shown that each principle implies the other. To prove that the well-ordering principle implies mathematical induction, assume P(0) is true and \forall n \, (P(n) \to P(n+1)). Suppose, for contradiction, that there exists some nonnegative integer where P fails; let C = \{ n \in \mathbb{N} \mid \neg P(n) \} be the nonempty set of counterexamples. By the well-ordering principle, C has a least element s. Since P(0) holds, s \geq 1. Then s-1 \notin C, so P(s-1) is true, and by the inductive hypothesis, P(s) follows, contradicting s \in C. Thus, C is empty, and P(n) holds for all n. Conversely, to prove that mathematical induction implies the well-ordering principle, let S be any nonempty subset of the nonnegative integers. Define the property Q(m) to mean that every nonempty subset of \{0, 1, \dots, m\} has a least element. First, verify Q(0): the only possible nonempty subset is \{0\}, whose least element is 0. Now assume Q(k) holds for some k \geq 0; consider any nonempty subset T \subseteq \{0, 1, \dots, k+1\}. If T \cap \{0, 1, \dots, k\} \neq \emptyset, then by Q(k) this intersection has a least element, which is also the least in T. Otherwise, T = \{k+1\}, so k+1 is the least element. Thus, Q(k+1) holds. By mathematical induction, Q(m) is true for all m. Since S is nonempty, let m \in S; then S \cap \{0, 1, \dots, m\} \neq \emptyset, so by Q(m) it has a least element, which is the least element of S.

Relation to the Axiom of Choice

The well-ordering principle, as applied to the natural numbers, asserts that every nonempty subset of the natural numbers has a least element and is provable in basic arithmetic without invoking the axiom of choice (AC). However, its generalization to arbitrary sets—the well-ordering theorem, stating that every set can be well-ordered—is equivalent to AC within Zermelo-Fraenkel set theory (ZF). This equivalence highlights a profound connection: while the principle holds unconditionally for countable well-ordered structures like the naturals, extending it universally requires the non-constructive selection mechanism provided by AC. In 1904, Ernst Zermelo published a proof that AC implies the well-ordering theorem, demonstrating that for any set, one can construct a well-ordering by iteratively selecting elements via a choice function on the power set. Zermelo's argument, motivated by David Hilbert's challenge to well-order the real numbers, introduced AC explicitly as a "logical principle" to justify the existence of such selections, sparking significant debate in the mathematical community about its intuitive validity. This historical development marked AC's formal debut, with Zermelo's theorem serving as its first major application, showing how choice enables well-orderings even for uncountable sets like the continuum. The converse implication—that the well-ordering theorem entails AC—holds in ZF set theory. Given a family of nonempty disjoint sets, well-order their union; then, for each set in the family, select the minimal element under this ordering to define a choice function. This construction relies on the theorem's guarantee of a total well-ordering, ensuring consistent minimal selections without further assumptions. Thus, in the combined framework of ZF plus AC (known as ), the well-ordering theorem is a core result, underscoring AC's role in extending the natural numbers' ordering principle to the full universe of sets.

Implications and Extensions

Implication from Completeness of the Real Numbers

The completeness axiom of the real numbers states that every nonempty subset of \mathbb{R} that is bounded above has a least upper bound (supremum) in \mathbb{R}. This property characterizes \mathbb{R} as a complete ordered field, distinguishing it from the rationals \mathbb{Q}, which lack completeness. This axiom implies the well-ordering principle for the natural numbers \mathbb{N} (under the standard ordering), where every nonempty subset of \mathbb{N} has a least element. To derive this, embed \mathbb{N} into \mathbb{R} via the natural inclusion map. Consider any nonempty subset S \subseteq \mathbb{N}. Since elements of S are nonnegative, S is bounded below by 0 in \mathbb{R}, and thus -S = \{-s \mid s \in S\} is bounded above. By the completeness axiom, \sup(-S) exists in \mathbb{R}, so \inf(S) = -\sup(-S) exists in \mathbb{R}. Let m = \inf(S). Since m is the greatest lower bound, m + 1 is not a lower bound, so there exists some n \in S with m \leq n < m + 1. The open interval (m, n) contains no elements of \mathbb{N} due to the discrete nature of \mathbb{N} in \mathbb{R}, implying that no element of S is less than n. Thus, n = m \in S is the least element of S. This proof for \mathbb{N} relies solely on the order completeness of \mathbb{R} and avoids the axiom of choice or transfinite methods.

Well-ordering in Other Ordered Sets

A partially ordered set (poset) is well-ordered if it is totally ordered and every non-empty subset has a least element. Finite sets equipped with their natural order are always well-ordered, as any non-empty subset is finite and thus possesses a minimal element. The set of natural numbers, denoted \omega, under the standard ordering forms the smallest infinite well-ordered set, where every non-empty subset has a least element, enabling principles like mathematical induction. Ordinal numbers, defined in set theory as the order types of well-ordered sets, are inherently well-ordered under the relation of proper initial segment (or membership), even in Zermelo-Fraenkel set theory without the axiom of choice. For instance, transfinite ordinals like \omega + 1 extend the ordering of \omega while preserving the well-ordering property. The rational numbers \mathbb{Q} are not well-ordered under their standard ordering, as subsets like the open interval (0, 1) \cap \mathbb{Q} lack a least element due to the density of \mathbb{Q}. However, since \mathbb{Q} is countably infinite, it admits a well-ordering via an explicit enumeration that embeds it order-isomorphically into \omega, without requiring the axiom of choice. The integers \mathbb{Z} under the standard ordering are not well-ordered, as the subset of negative integers has no least element. Nonetheless, \mathbb{Z} can be well-ordered constructively, for example, by the ordering $0, 1, -1, 2, -2, \dots, which yields an order type of \omega. The real numbers \mathbb{R} fail to be well-ordered under their standard ordering, since dense intervals like (0, 1) contain no least element. Moreover, any well-ordering of \mathbb{R} requires the axiom of choice, as no explicit such ordering is known and the well-ordering theorem equates the existence of well-orderings for all sets with choice.

Applications in Proofs

Proofs by Minimal Counterexample

Proofs by minimal counterexample constitute a proof technique in discrete mathematics that leverages the well-ordering principle to verify universal statements over the natural numbers. To establish that a property P(n) holds for every natural number n, the method assumes for contradiction that there exists some k where \neg P(k). The well-ordering principle is then invoked to identify the smallest such counterexample, allowing the derivation of an inconsistency that negates the assumption. This approach transforms the problem into analyzing a purported minimal violation, often revealing that the property must hold at that point or implying an even smaller breach. The reliance on the well-ordering principle is fundamental, as it ensures that the set of counterexamples, if non-empty, admits a least element among the nonnegative integers. The principle states that every non-empty subset of the natural numbers \mathbb{N} (including 0) has a minimal member under the standard ordering. Without this guarantee, the method could not proceed to isolate a "smallest" counterexample for scrutiny, which is essential for constructing the contradiction. This makes the technique particularly suited to domains where the natural numbers' ordered structure can be exploited to "peel away" layers of the problem. In practice, the general steps of a proof by minimal counterexample are structured as follows. Define the counterexample set C = \{ n \in \mathbb{N} \mid \neg P(n) \}. Suppose C is non-empty; by the , let m = \min(C), the smallest element in C. To reach a contradiction, show that P(m) actually holds—perhaps by relating m to smaller values where the property is known or assumed to be true—or demonstrate that the assumption \neg P(m) implies the existence of some k < m with \neg P(k), violating the minimality of m. Thus, C must be empty, proving P(n) for all n \in \mathbb{N}. This method offers advantages in its non-constructive nature, emphasizing the absence of counterexamples rather than explicit construction, which proves effective for establishing existence or universality in and algorithm correctness without requiring incremental verification. It provides a structured path to contradictions via the least counterexample, streamlining proofs in contexts where direct might be cumbersome.

Example: Absence of Integers Between 0 and 1

The well-ordering principle provides a straightforward way to demonstrate the absence of integers in the open interval between 0 and 1, underscoring the discrete structure of the integers. Specifically, the theorem states that there is no integer n satisfying $0 < n < 1. This result follows from the minimal counterexample approach, where the principle guarantees a least element in any nonempty subset of the positive integers. To prove the theorem, suppose for contradiction that the set S = \{ n \in \mathbb{Z} \mid 0 < n < 1 \} is nonempty. As S consists entirely of positive integers, the well-ordering principle implies that S possesses a minimal element m. However, every positive integer satisfies m \geq 1, which directly contradicts the condition m < 1. Thus, the assumption that S is nonempty must be false, so no such integer exists. This proof extends naturally to the natural numbers \mathbb{N} = \{0, 1, 2, \dots\}, where the set of natural numbers strictly between 0 and 1 is also empty by the same reasoning applied to the positive subset. The argument emphasizes the inherent discreteness of \mathbb{N} and \mathbb{Z}, with successive elements separated by gaps that contain no other members of the set. Overall, this example illustrates the well-ordering principle's fundamental role in establishing the ordered structure of the integers, confirming their lack of density in intervals like (0, 1) and reinforcing foundational properties used in more advanced proofs.

Example: Finiteness of Decreasing Nonnegative Integer Sequences

One key application of the well-ordering principle demonstrates the finiteness of strictly decreasing sequences in the nonnegative integers. Specifically, the theorem states that there does not exist an infinite strictly decreasing sequence in \mathbb{N}_0 = \{0, 1, 2, \dots \}. This follows directly from the principle that every nonempty subset of \mathbb{N}_0 possesses a least element. To prove this theorem using a minimal counterexample approach, assume for contradiction that such an infinite sequence exists: a_1 > a_2 > a_3 > \dots , where each a_i is a nonnegative integer. Consider the nonempty set S = \{a_i \mid i = 1, 2, 3, \dots \}. By the well-ordering principle, S has a minimal element, denoted a_k for some finite index k. However, the strict decrease implies a_{k+1} < a_k and a_{k+1} \in S, contradicting the minimality of a_k. Thus, the assumption is false, and every strictly decreasing sequence of nonnegative integers must terminate after finitely many terms. This theorem highlights the absence of infinite descending chains in well-ordered sets, a defining characteristic that ensures no endless strictly decreasing subsequences can occur. In general, a linearly ordered set is well-ordered if and only if it contains no infinite descending chain. The property extends beyond \mathbb{N}_0 to any well-ordered set, where the existence of minimal elements in every nonempty subset precludes infinite strict descents.

Example: Unique Prime Factorization

The Fundamental Theorem of Arithmetic asserts that every integer n > 1 can be expressed as a product of prime numbers in essentially one way, meaning that the prime factorization n = p_1^{e_1} p_2^{e_2} \cdots p_k^{e_k}, where the p_i are distinct primes and the e_i are positive integers, is unique up to the order of the factors. The existence of such a prime for every greater than 1 can be established using on n. To prove uniqueness, suppose toward a contradiction that there exists some m > 1 with two distinct prime factorizations, say m = p_1^{e_1} \cdots p_r^{e_r} = q_1^{f_1} \cdots q_s^{f_s}, where the p_i and q_j are primes (not necessarily distinct across lists) and the multisets \{p_1^{e_1}, \dots, p_r^{e_r}\} and \{q_1^{f_1}, \dots, q_s^{f_s}\} differ. Let S be the nonempty set of all such counterexamples m > 1; by the well-ordering principle, S has a least element m. , assume the primes in the first factorization are ordered so that p_1 is the smallest prime dividing m, and that p_1 does not equal any q_j with exponent at least as high as in the first factorization (after normalizing by dividing out the minimum powers of common primes from both sides). Then p_1 must divide the second factorization, so p_1 = q_k for some k, but adjusting exponents leads to p_1 dividing a proper of m obtained by dividing both sides by p_1^{e_1}, yielding a smaller (less than m) with two distinct prime factorizations, contradicting the minimality of m. Thus, no such m exists, proving uniqueness.

Advanced Properties

Non-algebraic Nature

The non-algebraic nature of the well-ordering principle refers to its independence from purely algebraic foundations, meaning it cannot be derived solely from the field axioms governing addition and multiplication in structures like the real or rational numbers. The field axioms ensure properties such as commutativity, associativity, and distributivity, but they do not guarantee the existence of a least element in every nonempty subset of positive elements, as this requires additional order or completeness assumptions. However, for the natural numbers, the well-ordering principle can be derived from the , as it is equivalent to the axiom within that system. While define the natural numbers through successor and , extending well-ordering to arbitrary sets demands axioms beyond finite arithmetic, such as the involving transfinite constructions. In , the well-ordering principle is not an intrinsic feature of operations; instead, analytic or set-theoretic tools are used to resolve existential claims. For instance, proving the existence of an for a —ensuring every has roots in some extension—invokes the (via the ) or to select bases or maximal chain extensions, but this step introduces non-constructive set-theoretic elements beyond pure algebra. Without such input, algebraic structures alone cannot enforce the global minimality required for well-ordering, distinguishing it from local properties like ring homomorphisms or ideal structures. In contrast to algebraic closures under operations like and , which follow deductively from axioms with , well-ordering pertains to holistic properties that cannot be captured by equation-solving or polynomial identities alone, emphasizing its reliance on choice-based orderings.

Role in Set Theory and Ordinals

In , ordinal numbers serve as a fundamental tool for measuring the order types of well-ordered sets, where two well-ordered sets are considered equivalent if there is an order-isomorphism between them. This partitions the class of all well-ordered sets into ordinal numbers, each representing a unique . The smallest ordinal, denoted ω, corresponds to the order type of the natural numbers ℕ equipped with the standard less-than ordering, providing the prototypical example of a countably well-ordering. A concrete realization of ordinals within is given by the von Neumann ordinals, introduced by in 1923. In this construction, each ordinal α is defined as the set of all ordinals β strictly less than α, i.e., α = {β | β < α}, with the ∈ serving as the ordering. This representation ensures that ordinals are transitive sets—every element is a —and that the class of all ordinals is well-ordered by ∈, inheriting the well-ordering property directly. The von Neumann thus embeds the entire structure of ordinals into the universe of sets, facilitating rigorous proofs about infinite structures without relying on external notions of order. Transfinite induction extends the principle of to the ordinals, leveraging their well-ordering to prove properties across transfinite sequences. Specifically, to establish that a property P holds for every ordinal α, it suffices to verify the induction step: assuming P(β) for all β < α, prove P(α). The base case P() follows from the being the zero ordinal. This method is valid because any non-empty class of ordinals has a least element, preventing infinite descending chains and ensuring the induction covers all ordinals exhaustively. is indispensable for defining operations like ordinal and recursively along the well-ordered class of ordinals. Key applications of well-ordering in include Hartogs' theorem, proved by Friedrich Hartogs in 1915, which asserts that for any set X, there exists a least ordinal α such that no well-ordering of type α injects into the power set 𝒫(X). This ordinal, known as the Hartogs number of X, is always a and guarantees the existence of arbitrarily large numbers (initial ordinals), which are the cardinalities of well-ordered infinite sets. The alephs, denoted ℵ_γ for ordinals γ, form the backbone of transfinite arithmetic and are constructed via transfinite recursion on ordinals. In advanced contexts like forcing, well-orderings underpin the construction of generic extensions by defining partial orders that preserve or alter cardinalities while maintaining well-foundedness; for instance, forcing posets often induce well-orderings on names for sets to ensure definability in the extension. Similarly, in the study of large cardinals, well-ordering is central, as these are typically initial ordinals κ with embedding properties (e.g., measurable cardinals admit ultrafilter measures), providing consistency strength hierarchies ordered by their ordinal magnitudes.

References

  1. [1]
    Well Ordering Principle -- from Wolfram MathWorld
    Well Ordering Principle: Every nonempty set of positive integers contains a smallest member. See also Axiom of Choice, Well Ordered Set.
  2. [2]
    [PDF] Well Ordering Principle: Chapter 2.1 – 2.3 - MIT OpenCourseWare
    Every nonempty set of nonnegative integers has a smallest element. This statement is known as The Well Ordering Principle. Do you believe it? Seems sort of ...
  3. [3]
    Set Theory - Stanford Encyclopedia of Philosophy
    Oct 8, 2014 · By the AC (in the form of the Well-Ordering Principle), every set \(A\) can be well-ordered, hence it is bijectable with a unique cardinal, ...2. The Axioms Of Set Theory · 6. The Set Theory Of The... · 10. Large Cardinals
  4. [4]
    [PDF] Induction - UNL School of Computing
    At its heart is the Well Ordering Principle. Theorem (Principle of Well Ordering) Every nonempty set of nonnegative integers has a least element. Since every ...
  5. [5]
    [PDF] Induction and the Well Ordering Principle - People | MIT CSAIL
    3.1 The Well Ordering Principle. Every nonempty set of nonnegative integers has a smallest element. This statement is known as The Well Ordering Principle.
  6. [6]
    Formal statement of well-ordering (not the theorem)
    Oct 31, 2015 · Let (S,⪯) be an ordered set. Then the ordering ⪯ is a well-ordering on S iff every non-empty subset of S has a smallest element under ⪯: ∀T ...Use the Well ordering principle to prove - Math Stack Exchangeelementary set theory - Principle of Induction and Well-Ordering ...More results from math.stackexchange.com
  7. [7]
    [PDF] Partial Order, Total Order and Well-ordering - 4dspace@MTTS
    We say that the order ≤ is a well-ordering on X if it is (i) a total order and (ii) every nonempty subset of X has a minimum. 37. Well-ordering Principle. On ...
  8. [8]
    [PDF] Lecture 3 1 Overview 2 Well-Ordering Principle
    Jan 16, 2019 · The well-ordering principle (WOP) states that there is a smallest number in any set of positive integers, and the set of all positive integers ...
  9. [9]
    On the origin of the well-ordering principle of $\mathbb{Z}^{+}
    Jan 17, 2025 · E. Zermelo, 1904, spoke of well ordering as a principle ("Prinzip") in Beweis, daß jede Menge wohlgeordnet werden kann (Proof that every set ...Missing: Bolzano | Show results with:Bolzano
  10. [10]
    [PDF] Notes on Richard Dedekind's Was sind und was sollen die Zahlen?
    Notes on Richard Dedekind's. Was sind und was sollen die Zahlen? David E ... Dedekind says that the preceding theorem is the basis for mathematical induction, the.
  11. [11]
    Zermelo's Axiomatization of Set Theory
    Jul 2, 2013 · 2.2.2 Zermelo's 1904 Proof of the Well-Ordering Theorem. Zermelo's approach to the well-ordering problem took place in three stages. He ...The Background to Zermelo's... · The Well-Ordering Problem... · Completeness
  12. [12]
    [PDF] Induction and the Well Ordering Principle - People | MIT CSAIL
    equivalent in the sense that whatever you can prove using one of the methods, you can also prove using either of the others. The choice of which method to ...
  13. [13]
    [PDF] Math 310 Class Notes 4: The Well-ordering Principle
    The well-ordering principle states that every nonempty subset T of N has a least element, where m ≤ n for all n ∈ T.Missing: formal | Show results with:formal<|control11|><|separator|>
  14. [14]
    The Axiom of Choice - Stanford Encyclopedia of Philosophy
    Jan 8, 2008 · Zermelo introduces axioms of set theory, explicitly formulates AC and uses it to prove the well-ordering theorem, thereby raising a storm of ...
  15. [15]
    None
    Below is a merged summary of the Well-Ordering Principle and Completeness of Real Numbers in Rudin's *Principles of Mathematical Analysis*, consolidating all information from the provided segments into a dense, structured format. Given the request for a "dense representation" and the potential utility of tables, I will use a combination of text and CSV-style tables to retain all details efficiently. The response avoids any "thinking tokens" beyond the final output, focusing solely on the synthesis of the provided summaries.
  16. [16]
    [PDF] Chapter 8 Ordered Sets
    a least upper bound , then the least upper bound is unique (. ). B why? vi ... 3) Order isomorphisms preserve well-ordering: if a poset is well-ordered ...
  17. [17]
    [PDF] The Axiom of Choice and Some Equivalences: - Kenyon College
    Nov 29, 2012 · Following this we will show in a round- about, yet elegant way, the equivalence of the Axiom of Choice to the Well-. Ordering Principle, Zorn's ...
  18. [18]
    [PDF] Minimal Counterexample: A Different look at Induction1
    Aug 28, 2021 · What we have used before, implicitly and rather matter-of-fact-ly, is the following axiom called the well- ordering principle (WOP). Any non- ...
  19. [19]
    Strong Induction and Well-Ordering | Discrete Mathematics Class ...
    Proof by Minimum Counterexample. Combines the Well-Ordering Principle with proof by contradiction; Steps in a Proof by Minimum Counterexample: ... Advantages and ...
  20. [20]
    [PDF] Number Theory for Mathematical Contests - UTRGV Faculty Web
    Apr 1, 2016 · ... no integer between 0 and 1. 2 Example Prove that there is no ... . Proof: We use the Well-Ordering Principle. Consider the set S = {a ...
  21. [21]
    [PDF] SEQUENCES, MATHEMATICAL INDUCTION, AND RECURSION ...
    Another consequence of the well-ordering principle is the fact that any strictly decreasing sequence of nonnegative integers is finite. That is, if r. 1. , r. 2.Missing: finiteness | Show results with:finiteness
  22. [22]
    Greatest Common Divisors
    for some n. Note that $a_1 > a_2 > a_3 > \cdots$ is a decreasing sequence of nonnegative integers. The well-ordering principle implies that this sequence cannot ...
  23. [23]
    [PDF] Untitled - UCLA Department of Mathematics
    A linear ordering (A,≤) is a wellordering if and only if there is no infinite descending chain x0 > x1 > ททท. Hint: This requires a mild form of the Axiom ...
  24. [24]
    [PDF] The Fundamental Theorem of Arithmetic
    Both parts of the proof will use the. Well-ordering Principle for the set of natural numbers. (1) We first prove that every a > 1 can be written as a product of ...
  25. [25]
    [PDF] 2.2 The Fundamental Theorem of Arithmetic
    By the Well-Ordering Principle, there is a smallest such natural number. Let N be this smallest natural number. Thus, if 1 <n<N, then n can be ex- pressed as ...
  26. [26]
    [PDF] Divisibility, Factoring, Primes. The Fundamental Theorem of Arithmetic.
    May 29, 2008 · The Division Algorithm and the Euclidean Algorithm. The Well Ordering Prin- ciple is the main ingredient in the proof of the following statement ...
  27. [27]
    [PDF] Chapter 4 - MAT246H1S Lec0101 Burbulla
    4.1: Proof of The Fundamental Theorem of Arithmetic. Chapter 4 Lecture Notes ... By the well-ordering principle there is a least number in T, call it N ...
  28. [28]
    LPT The Proof of the Fundamental Theorem of Arithmetic
    By the Principle of Well-Ordering, \(S\) has a smallest number, say \(a\text{.}\) If the only positive factors of \(a\) are \(a\) and \(1\text{,}\) then \(a ...
  29. [29]
    [PDF] Chapter 5 Three Famous Theorems
    By the Well-Ordering Principle (Theorem 4.34), S contains a least element, say n. Then n cannot be prime since this would satisfy the theorem. So, it must be ...<|control11|><|separator|>
  30. [30]
    [PDF] 21-355 Principles of Real Analysis I Fall 2004 I. The Real Number ...
    In analysis textbooks this is often referred to as the well-ordering principle (for the natural natural numbers). Some people refer to it as the least number ...
  31. [31]
    [PDF] Zorn's lemma and some applications, II - Keith Conrad
    Steinitz was hindered in this work by the primitive state of set theory at that time and he used the well-ordering principle ... dealing with non-algebraic ...<|separator|>
  32. [32]
    Hilbert's Program - Stanford Encyclopedia of Philosophy
    Jul 31, 2003 · It calls for a formalization of all of mathematics in axiomatic form, together with a proof that this axiomatization of mathematics is consistent.Historical development of... · Hilbert's Program and Gödel's...
  33. [33]
    [PDF] Mathematics 4530 Ordinal numbers and the well-ordering theorem
    Well-orderings are useful because they allow proofs by induction: Induction principle. Let X be a well-ordered set and let Y be a subset. Suppose that for all x ...
  34. [34]
    [PDF] Transfinite Induction - Penn Math
    Finally ordinals: the intuitive definition is that an ordinal number is a well ordered order type, that is, an equivalence class of well orders. Although there ...
  35. [35]
    [PDF] The Ordinal Numbers and Transfinite Induction - Purdue Math
    Sep 14, 2015 · An ordinal number can be thought of as the position of an element in a well-ordered set. Example. Let N ∪ {ω} have the same ordering as before.