Fact-checked by Grok 2 weeks ago

Lagrange's theorem

Lagrange's theorem is a fundamental result in group theory stating that if G is a and H is a of G, then the order of H, denoted |H|, divides the order of G, denoted |G|. This theorem establishes a key divisibility property that constrains the possible structures of finite groups and their subgroups. Named after the mathematician , the theorem originated in his 1770 work Réflexions sur la résolution algébrique des équations, where he examined of variables in polynomial equations and observed that the number of distinct polynomials obtained divides n! for an equation of degree n. Although Lagrange did not frame his result in terms of abstract groups—which were not formalized until the —his insights laid foundational groundwork for theory. In 1815 and 1844, extended Lagrange's ideas by proving related results for , showing that the order of a divides n! in the S_n. Camille Jordan further generalized the theorem to all finite in his 1861 doctoral thesis, and by the 1880s, with the emergence of abstract , Otto Hölder and Heinrich Weber provided proofs using cosets that align with the modern statement. The theorem's proof relies on partitioning the group G into disjoint left cosets of H, each of the same cardinality as H, leading to |G| = [G : H] \cdot |H|, where [G : H] is the index of H in G. Notable corollaries include that every group of prime order is cyclic, since the only subgroups are the trivial group and itself, and that the order of any element in G divides |G|, as the subgroup generated by an element has order equal to the element's order. These implications are pivotal in classifying finite groups and understanding their symmetries. Beyond pure , Lagrange's theorem has significant applications in ; for instance, it provides a group-theoretic proof of by considering the of integers modulo a prime, where the order of any element divides p-1. Similarly, it underpins in the context of the modulo n. The theorem also informs modern fields like , where properties of cyclic groups of prime order ensure security in protocols such as Diffie-Hellman .

Statement and Prerequisites

Formal Statement

Lagrange's theorem, a fundamental result in group theory, asserts that if G is a and H is a of G, then the order of H, denoted |H|, divides the order of G, denoted |G|. Equivalently, |G| = [G : H] \cdot |H|, where [G : H] denotes the index of H in G, defined as the number of distinct left cosets (or equivalently, right cosets) of H in G. This statement extends to infinite groups via cardinal arithmetic: if G is any group and H a subgroup, then the cardinality of G equals the product of the cardinality of H and the cardinality of the set of left cosets of H in G. The theorem is named after the Italian-French mathematician (1736–1813), who developed related ideas in his 1770 work Réflexions sur la résolution algébrique des équations, focusing on permutations of roots in polynomial equations, though he did not formulate it in the abstract group-theoretic sense. It should be distinguished from other results bearing Lagrange's name, such as the in analysis or Lagrange's theorem in .

Essential Definitions

A group G is a set equipped with a binary operation, often denoted by multiplication or addition, that satisfies four fundamental axioms: closure, under which the product of any two elements in G remains in G; associativity, under which the grouping of operations does not matter; the existence of an identity element e such that e \cdot g = g \cdot e = g for all g \in G; and the existence of inverses, under which every element g has an element g^{-1} satisfying g \cdot g^{-1} = g^{-1} \cdot g = e. These axioms ensure that the structure captures the essential properties of symmetry and transformation. While groups can be infinite, the finite case is particularly relevant for many applications, including Lagrange's theorem. A H of a group G is a nonempty of G that itself forms a group under the same restricted to H, meaning it satisfies , associativity (inherited from G), contains the of G, and includes inverses for each of its elements (which are also in G). Subgroups provide building blocks for understanding the structure of larger groups. The of a group G, denoted |G|, is the of the set G, representing the number of distinct elements in G. For finite groups, this is a positive ; for infinite groups, it is infinite. The finite order is central to results like Lagrange's theorem, which connects the orders of a group and its subgroups. Given a subgroup H of a group G, a left coset of H is a set of the form gH = \{ gh \mid h \in H \} for some g \in G, while a right coset is Hg = \{ hg \mid h \in H \}. Each coset has the same number of elements as H, and the collection of left (or right) cosets partitions G. Left and right cosets coincide for every element g \in G if and only if H is a of G, though the full implications of normality are explored elsewhere. The index of a subgroup H in a group G, denoted [G : H], is the number of distinct left cosets of H in G. This measure quantifies how H "sits inside" G, and for finite groups, it equals |G| / |H| by Lagrange's theorem.

Proof and Extensions

Standard Proof

The standard proof of Lagrange's theorem relies on the partition of a group G into left cosets of a subgroup H. For a finite group G and subgroup H \leq G, the left cosets gH = \{gh \mid h \in H\} for g \in G form a partition of G into disjoint subsets, where the number of distinct cosets is the index [G:H]. Each left coset gH has exactly |H| elements, as the map \phi: H \to gH defined by \phi(h) = gh is a : it is injective because if gh_1 = gh_2, then h_1 = h_2 by left cancellation in G; it is surjective by construction. The s are pairwise disjoint: if g_1 H \cap g_2 H \neq \emptyset, then there exist h_1, h_2 \in H such that g_1 h_1 = g_2 h_2, so g_1^{-1} g_2 = h_1 h_2^{-1} \in H, which implies g_2 H = g_1 (g_1^{-1} g_2) H = g_1 H. Thus, G is the disjoint union of [G:H] cosets, each of cardinality |H|, so |G| = [G:H] \cdot |H|. It follows that |H| divides |G|. For the infinite case, the argument generalizes using cardinal arithmetic: the cardinality of G equals the cardinal product of the index [G:H] (possibly infinite) and |H|, so |H| divides |G| in the sense of cardinal division. An alternative perspective views Lagrange's theorem as a consequence of the orbit-stabilizer theorem in group actions; specifically, the action of G on the set of left cosets of H by left has orbits of size [G:H] and stabilizers isomorphic to H, linking the directly to group orders.

Chain Rule Extension

A fundamental extension of Lagrange's theorem concerns chains of subgroups, where the exhibits a multiplicative property. Specifically, if K \leq H \leq G are subgroups of a G, then the [G : K] = [G : H] \cdot [H : K]. To prove this, consider the cosets of H in G, of which there are [G : H], partitioning G into disjoint sets each of cardinality |H|. Within each such coset gH, the cosets of K in H refine it further, since K \leq H, yielding [H : K] cosets of K per coset of H. Thus, the total number of cosets of K in G is the product [G : H] \cdot [H : K]. This refinement establishes a between the of K in G and the pairs of (gH, hK) with g \in G and h \in H: the map sending ghK to (gH, hK) is well-defined, injective, and surjective, as distinct pairs yield distinct and every arises this way. As an immediate , combining with Lagrange's theorem implies that |K| divides |H|, which in turn divides |G|, providing a chain of divisibility for orders in finite groups.

Applications

Consequences for Elements and Orders

One immediate consequence of Lagrange's theorem is that the order of any element in a finite group divides the order of the group. If g \in G and the order of g is n, then the cyclic subgroup \langle g \rangle generated by g has order n, so n divides |G| by the theorem. This implies that g^{|G|} = e, the identity element, for every g \in G. Lagrange's theorem also highlights the structure of trivial subgroups. The trivial subgroup \{e\} has 1, which divides |G| for any G. Similarly, the full group G itself is a of |G|, with 1 in G. These cases illustrate the theorem's boundary conditions, ensuring that subgroup orders are consistent with divisibility even in the most basic scenarios. When |G| = p for a prime p, Lagrange's theorem yields strong structural results. Any non-identity element g \in G has order dividing p, so its order must be p (since it cannot be 1). Thus, \langle g \rangle = G, making G cyclic and generated by any such g. Moreover, G has no proper nontrivial subgroups, as their orders would divide p but exceed 1, which is impossible. This absence of proper nontrivial subgroups implies that G is simple, with no nontrivial normal subgroups. Sylow's theorems extend the implications of Lagrange's theorem specifically for prime power orders. For a finite group G with |G| = p^k m where p is prime and \gcd(p, m) = 1, the first Sylow theorem guarantees the existence of a Sylow p-subgroup of order p^k, the highest power of p dividing |G|. This provides a partial converse to Lagrange's theorem by ensuring subgroups exist for certain divisor orders that the theorem alone does not predict. The remaining Sylow theorems describe the number and conjugacy of these subgroups, further constraining group structure. In modern group theory, these consequences underpin the classification of finite simple groups. Lagrange's theorem, through its role in dictating possible orders and enabling tools like , imposes critical constraints on the orders and structures of simple groups, facilitating their exhaustive enumeration.

Connections to Number Theory

Lagrange's theorem finds profound applications in through the structure of multiplicative groups of units modulo an integer. For a prime p, the group (\mathbb{Z}/p\mathbb{Z})^* consists of the integers from 1 to p-1 under modulo p, and it has p-1. By Lagrange's theorem, the of any divides the group ; in particular, for any a not divisible by p, the cyclic generated by a has dividing p-1, so a^{p-1} \equiv 1 \pmod{p}. This is , a cornerstone of . Euler's theorem generalizes this result to composite moduli. The group (\mathbb{Z}/n\mathbb{Z})^* has order \phi(n), where \phi is Euler's totient function counting integers up to n coprime to n. Thus, for \gcd(a, n) = 1, the order of a in this group divides \phi(n), yielding a^{\phi(n)} \equiv 1 \pmod{n}. Both Fermat's and Euler's theorems follow directly from applying Lagrange's theorem to the cyclic subgroup \langle a \rangle generated by such an a, as the finite order of the group ensures the exponent of a divides |(\mathbb{Z}/n\mathbb{Z})^*|. In modern , Lagrange's theorem underpins the security of protocols relying on finite abelian groups. For instance, in elliptic curve Diffie-Hellman , the points on an over a form a group whose N is known; Lagrange's theorem guarantees that the of any or divides N, allowing selection of large prime-order subgroups to resist attacks while ensuring computational feasibility. This property is essential for the efficiency and security of standards. Similarly, in coding theory, Lagrange's theorem informs the design of error-correcting codes over finite fields. Reed-Solomon codes, defined over \mathbb{F}_q, evaluate polynomials at points from a multiplicative subgroup of \mathbb{F}_q^*, which has order q-1; the theorem ensures that element orders divide q-1, enabling the choice of primitive elements to generate full cycles for evaluation points and consecutive roots in the BCH bound, thus achieving optimal minimum distance for error detection and correction. Lagrange's theorem also connects to classical results like , proved via the pairing of inverses in (\mathbb{Z}/p\mathbb{Z})^* for odd prime p, where the group's finite order p-1 facilitates showing the product of units is -1 \pmod{p}. The converse of —that p prime iff (p-1)! \equiv -1 \pmod{p}—implies infinitely many primes: for any integer n > 1, (n-1)! + 1 has a prime factor r \geq n, as no prime less than n divides it (else it would divide 1), so repeated application yields arbitrarily large primes.

Limitations and Converse

Failure of the Converse

The converse of Lagrange's theorem asserts that if d divides the of a G, then G possesses a of d. This statement fails to hold in general for arbitrary . Certain cases guarantee the existence of subgroups for specific divisors of |G|. Every G contains of 1 (the trivial subgroup) and |G| (itself). Moreover, by Cauchy's theorem, if a prime p divides |G|, then G contains an element of p, and hence a of p. A G is termed a CLT-group if it satisfies the converse of Lagrange's theorem for every of |G|. Cyclic groups provide a fundamental class of CLT-groups, as each of their corresponds to a unique cyclic . Similarly, finite p-groups are CLT-groups, possessing subgroups of every dividing their p-power via iterative central series constructions. In the broader context of solvable groups, Hall's theorem ensures the existence of Hall \pi- for any set of primes \pi dividing |G|, where such a has order divisible only by primes in \pi and index coprime to that order—thus guaranteeing of orders dividing |G| precisely when the order and its complementary index are coprime. Burnside's p-complement theorem offers a conditional extension: if P is a Sylow p- of a G contained in of its normalizer N_G(P), then G admits a p-complement, a of index |P|.

Key Counterexamples

A classic example illustrating that the converse of Lagrange's theorem can hold is the S_3, which has 6 and possesses subgroups of every dividing 6, namely orders 1, 2, 3, and 6. Specifically, S_3 contains trivial subgroups of 1, cyclic subgroups of 2 generated by transpositions, the alternating subgroup A_3 of 3, and itself as the subgroup of 6. In contrast, the alternating group A_4 provides a key , as it has order 12 but no of order 6. The group A_4 consists of all even permutations of four elements, and its order is \frac{4!}{2} = 12. To see why no such exists, suppose for that H is a of order 6. Then H cannot be cyclic, as A_4 has no element of order 6, so H \cong S_3. The group S_3 has two elements of order 3. However, A_4 contains eight 3-cycles (elements of order 3). Since [A_4 : H] = 2, H is in A_4. Any 3-cycle has order 3, which does not divide 2, so its image in the quotient A_4 / H \cong \mathbb{Z}_2 must be trivial, implying all eight 3-cycles lie in H. But H can contain at most two elements of order 3, a . Another prominent counterexample is the A_5, which has 60 but no of 30. Any of 30 would have index 2 and thus be , but A_5 is and admits no nontrivial proper subgroups. Similarly, A_5 lacks subgroups of 20, though the focus here remains on the index-2 case for 30.

Historical Development

Early Origins

The origins of Lagrange's theorem trace back to the 18th and early 19th centuries, emerging within the contexts of and the algebraic study of polynomial equations rather than in the framework of abstract , which would not be formalized until later. Mathematicians of this era investigated permutations of and , laying implicit groundwork for ideas about orders dividing group orders through concrete examples in symmetric and multiplicative structures. Joseph-Louis Lagrange's contributions in the 1770s arose from his efforts to resolve higher-degree equations by radicals, where he examined of roots and their effects on symmetric functions. In his 1770–1771 memoir Réflexions sur la résolution algébrique des équations, Lagrange demonstrated that if a function of n variables assumes r distinct values under all n! of its arguments, then r divides n!, an early result implicitly connecting sizes to the full order in the context of equation solvability. This work built on as tools for analyzing resolvents, without explicit group-theoretic language. Carl Friedrich Gauss advanced a specific case in 1801 through his seminal Disquisitiones Arithmeticae, proving that in the of integers a prime p, the of any divides p-1. In Article 55, Gauss showed that the minimal exponent e satisfying a^e ≡ 1 mod p for a nonzero a p divides p-1, establishing the for this structure central to number-theoretic congruences. Augustin-Louis Cauchy provided an early explicit result in his studies of . In his 1845 , Cauchy proved that if a prime p divides the of a finite acting on n letters, then the group contains a of p, yielding a cyclic of that — a precursor to broader divisibility properties in symmetric settings. Camille Jordan generalized these insights in 1861 with his doctoral thesis Mémoire sur le nombre des valeurs des fonctions développées sous forme de séries infinies, where he established the full for arbitrary finite permutation groups: the order of any divides the order of the group. This marked a pivotal step toward formulations, still confined to permutation representations.

Modern Formulations

The adoption of into marked a significant shift from its origins in permutation groups to a general principle applicable to any finite group structure. Camille Jordan provided an early generalization in 1861, extending the theorem to arbitrary finite permutation groups by showing that the order of any divides the order of the group through the of orbit-stabilizer relations. This was further formalized for groups by Heinrich Weber in his 1895–1896 textbook Lehrbuch der Algebra, where he introduced the modern decomposition proof, establishing that if H is a of a finite group G, then |G| = |H| \cdot [G:H], with [G:H] denoting the index of H in G. Subsequent texts, such as those by Otto Hölder in 1889 and Bartel van der Waerden in 1930, refined this using equivalence relations on , solidifying its place as a cornerstone of . In the 20th century, extensions of the theorem integrated it with representation theory and topological structures. In representation theory, a key corollary states that the degree of any irreducible complex character of a finite group G divides |G|, derived from the fact that the representation yields a projective module over the group algebra whose dimension satisfies the divisibility condition via Frobenius reciprocity and inner product orthogonality. This result, central to understanding group symmetries, appears prominently in I. Martin Isaacs' Character Theory of Finite Groups (Theorem 2.23). For topological groups, particularly compact Lie groups, an analogous formulation emerges through the Peter-Weyl theorem, where the degrees of irreducible unitary representations are finite integers, and the regular representation decomposes with multiplicities equal to the degrees, mirroring the finite case in the L² sense despite the infinite "order"; this ensures that representation dimensions align with the group's structure in a measure-theoretic index. To address computational gaps in verifying the theorem for large groups, 20th-century developments incorporated it into algorithmic frameworks. The Schreier-Sims algorithm, introduced by Sims in the 1970s and building on Otto Schreier's 1930s work, computes base-strong generating sets for permutation groups, enabling efficient coset enumeration and order calculation that directly applies Lagrange's divisibility to confirm subgroup indices without exhaustive enumeration. This has been implemented in systems like , facilitating practical checks for groups up to enormous orders. Recent advancements have generalized the theorem beyond traditional groups. In 2011, Marcelo Aguiar and Swapneel Mahajan proved a version for Hopf monoids in the category of , stating that the dimension of a sub-Hopf monoid divides the dimension of the ambient Hopf monoid, extending Radford's Hopf algebra analog via comodule decompositions. In categorical settings, analogs appear in representations, where orbit sizes divide total objects under finite actions. The theorem also played a foundational role in the , completed in 2004 by Michael Aschbacher and Stephen D. Smith, where divisibility constraints bounded possible orders and Sylow structures across the 26 sporadic groups and Lie-type families. For groups, the formulation shifts to arithmetic: if H \leq G, then |G| = |H| \cdot [G:H] in the sense of cardinal multiplication, holding under the and emphasizing that infinite subgroups may have the same as the group, unlike the finite case.

References

  1. [1]
    [PDF] Cosets and Lagrange's theorem - Keith Conrad
    Lagrange's theorem leads to group-theoretic explanations of some divisibility ... Now we turn to applications of Lagrange's theorem in group theory itself.
  2. [2]
    [PDF] Lagrange's Theorem: Statement and Proof - St. Olaf College
    Apr 5, 2002 · If G is a group with subgroup H, then the left coset relation, g1 ∼ g2 if and only if g1 ∗ H = g2 ∗ H is an equivalence relation. Proof ...
  3. [3]
    [PDF] The History of Lagrange's Theorem - maths.nuigalway.ie
    The initial work that Lagrange did on polynomials bore little resemblance to the theorem that we have today. Group theory was not defined at this point of time.
  4. [4]
    [PDF] Consequences of Lagrange's Theorem
    Let H Ă G be a subgroup of G. Then Lagrange's Theorem tells us that |H| divides |G|. Since |G| is prime, |H| must be 1 or |G|. In the first case, we must have ...
  5. [5]
    Lagrange's Theorem and its Applications in Group Theory
    The theorem states that if G is a finite group and H is a subgroup of G, then the order of H divides the order of G. This paper will first cover elementary ...Missing: explanation | Show results with:explanation
  6. [6]
    [PDF] Cyclic groups. Cosets. Lagrange's theorem.
    Theorem (Lagrange) If H is a subgroup of a finite group G, then o(G)=[G : H] · o(H). In particular, the order of H divides the order of G. Proof: For any a ∈ G ...Missing: explanation | Show results with:explanation
  7. [7]
    None
    Below is a merged response that consolidates all the information from the provided summaries into a single, comprehensive overview. To maximize density and clarity, I will use tables in CSV format where appropriate to organize the details efficiently. The response will cover the formal statements, historical notes, and useful URLs for each segment, ensuring all information is retained.
  8. [8]
    [PDF] ALGEBRA 1 FOR COMPUTER SCIENTISTS kompatscher@karlin ...
    We remark that its statement is even correct for infinite groups (using cardinal arithmetic, i.e. ... By Lagrange's theorem then ord(a) = |H| must divide ...<|control11|><|separator|>
  9. [9]
    [PDF] The History of Lagrange's Theorem - maths.nuigalway.ie
    Introduction. The Italian mathematician Joseph Louis Lagrange created a very famous theorem in groups and applied mathematics known as Lagrange's Theorem.
  10. [10]
    Group -- from Wolfram MathWorld
    A group is a monoid each of whose elements is invertible. A group must contain at least one element, with the unique (up to isomorphism) single-element group ...
  11. [11]
    Subgroup -- from Wolfram MathWorld
    A subgroup is a subset H of group elements of a group G that satisfies the four group requirements. It must therefore contain the identity element.Missing: definition | Show results with:definition
  12. [12]
    Group Order -- from Wolfram MathWorld
    ### Summary of Group Order from Wolfram MathWorld
  13. [13]
    Coset -- from Wolfram MathWorld
    A subset of G of the form xH for some x in G is said to be a left coset of H and a subset of the form Hx is said to be a right coset of H.
  14. [14]
    [PDF] §3.8 Cosets, Normal Subgroups, and Factor Groups
    Looking ahead: H is normal if and only if its left and right cosets coincide. In particular, for abelian groups, left cosets and right cosets are the same.
  15. [15]
    [PDF] 8. Lagranges Theorem Definition 8.1. Let G be a group and let H be ...
    Definition 8.1. Let G be a group and let H be a subgroup. The index of H in G, denoted [G : H], is equal to the number of left cosets of H in G.Missing: explanation | Show results with:explanation
  16. [16]
    [PDF] 1 Cosets 2 Lagrange's Theorem and Related Results
    Oct 12, 2016 · Orbit-Stabilizer Theorem (Thm. 7.3): |orbG(s)| = [G : stabG(s)]. • Orbit–Stabilizer Thm follows from: Lemma: For ϕ, ψ ∈ G, ϕ(s) = ψ(s) iff ...Missing: equivalent | Show results with:equivalent
  17. [17]
    [PDF] Algebra, Second Edition - CSE IITB
    ... Multiplicative Property of the Index. Let G. H K be subgroups of a group G. Then [G: K] = [G: H][H: K]. Proof We will assume that the two indices on the ...
  18. [18]
    Index is multiplicative - Groupprops
    Nov 15, 2015 · This article gives the statement, and possibly proof, of a subgroup property (ie, subgroup of finite index) satisfying a subgroup metaproperty.
  19. [19]
    [PDF] Group Theory Notes
    Mar 5, 2011 · Definition 2.8: The group of cosets of a normal subgroup N of the group G is called the quotient group or the factor group of G by N. This group ...
  20. [20]
    [PDF] Study Guide for Algebra - Amherst College
    That is, |H| |G|. When g is an element of a finite group G, Lagrange's Theorem has several important consequences: • If |G| is prime, ...
  21. [21]
    Sylow Theory
    The Sylow E-theorem can be viewed as a partial converse of Lagrange's theorem. Lagrange asserts that if H is a subgroup of G and |H| = k, then k divides |G|. ...
  22. [22]
    [PDF] Section 36 Sylow Theorems - Columbia Math Department
    Sylow's Theorem and Lagrange's Theorem are the two most important results in finite group theory. The first gives a sufficient condition for the existence of ...Missing: extension | Show results with:extension
  23. [23]
    [PDF] On the Classification of Finite Simple Groups - MIT Mathematics
    May 22, 2022 · Obser- vations of cosets help us to prove one of the most applicable theorems in group theory, Lagrange's Theorem. Theorem 2.10 (Lagrange).
  24. [24]
    Number Theory - The Order of a Unit
    Fermat's Little Theorem​​ Theorem: Let be a prime. Then a p = a ( mod p ) for any a ∈ Z p . This theorem is often equivalently stated as a p − 1 = 1 for nonzero ...
  25. [25]
    [PDF] the euler-fermat theorem and group theory - Daniel Mathews
    Returning to the proof of Lagrange's theorem, we see that the group G is divided into equivalence classes with n elements. Hence the total number of elements in.
  26. [26]
    [PDF] Elliptic Curve Cryptography - BSI
    Elliptic curve cryptography (ECC) is a very efficient technology to realise public key cryptosys- tems and public key infrastructures (PKI).<|separator|>
  27. [27]
    [PDF] Introduction to finite fields - Stanford University
    In the next chapter, finite fields will be used to develop Reed-Solomon (RS) codes, the most useful class of algebraic codes. Groups and polynomials provide ...
  28. [28]
    [PDF] FFermat, Euler, Wilson, Linear Congruences, Lecture 4 Notes
    Theorem 20 (Wilson's Theorem). If p is a prime then (p − 1)! ≡ −1 mod p. Proof. Assume that p is odd (trivial for p = 2). Lemma 21. The congruence x2 ≡ 1 ...
  29. [29]
    [PDF] A Collection of Proofs regarding the Infinitude of Primes
    Dec 14, 2013 · proof that there exists infinitely many primes had only the slightest practical importance. ... Another example is Wilson's theorem which ...
  30. [30]
    [PDF] Theory and Applications - Abstract Algebra
    Aug 5, 2017 · ... converse of Lagrange's Theorem is false). The group A4 has order 12 ... does not hold for Z[x]. Why does it fail? 14. Prove or disprove ...<|separator|>
  31. [31]
    [PDF] proof of cauchy's theorem - Keith Conrad
    (Cauchy) Let G be a finite group and p be a prime factor of |G|. Then G contains an element of order p. Equivalently, G contains a subgroup of order p. The ...Missing: citation | Show results with:citation
  32. [32]
    [PDF] CLT-groups with cyclic or abelian subgroups - arXiv
    Jun 17, 2025 · A finite group G, which satisfies the converse of Lagrange's theorem, is called a CLT-group. A natural number n is said to be a CLT number ...<|separator|>
  33. [33]
    [PDF] Hall's theorems on solvable groups - UChicago Math
    It states that Hall π-subgroups are conjugate and furthermore any π-subgroup is contained in a Hall π-subgroup. This result gives us a great level of ...
  34. [34]
    [PDF] Group Theory - Berkeley Math
    Burnside's normal p-complement states that if G is a finite group and if P is a Sylow p-subgroup of G and if CG(P) = NG(P) then G contains a normal p-complement ...Missing: source | Show results with:source
  35. [35]
    [PDF] The symmetric group
    Theorem 3.1 (Cayley's Theorem). Every group of order n is isomorphic to a subgroup of. Sn. Proof. Suppose G a group of order n. Let G operate on itself by ...
  36. [36]
    A<sub>4</sub> Definitely Has no Subgroup of Order Six! - jstor
    But V is a group of order 4 and 4 does not divide 6, contradicting Lagrange's theorem. Proof 8. (Using the commutator subgroup). Since H is a subgroup of index ...
  37. [37]
    [PDF] Abstract Algebra: Supplementary Lecture Notes
    Recall that the standard counterexample to the converse of Lagrange's theorem is the alternating group A4, which has 12 elements but no subgroup of order 6.
  38. [38]
    [PDF] Resonance July 2012 Cover Tp.cdr
    The standard example of the alternating group A4, which has order 12 but has no subgroup of order 6, shows that the converse of Lagrange's theorem is not true.
  39. [39]
    82.34 A Note on the Converse to Lagrange's Theorem - jstor
    The most cited counter-example is the group A4. (see below) which has order 12 but no subgroup of order 6. Unlike various proofs [1-4] which use cosets ...
  40. [40]
    [PDF] Simplicity of A5 - Columbia Math Department
    Apr 11, 2020 · The alternating group A5 is simple. Theorem. The alternating group A5 ⊂ S5 is a simple group of order 60. In fact we have the general theorem:.
  41. [41]
    [PDF] Converse Lagrange Theorem Orders and Supersolvable Orders
    Nov 13, 2016 · ... example, A4, the alternating group on four symbols, which has order 12, has no subgroup of order 6 [2]. If G is a group which has a subgroup ...
  42. [42]
    [PDF] A History of Lagrange's Theorem on Groups
    Lagrange's Theorem first appeared in 1770-71 in connection with the problem of solving the general polynomial of degree 5 or higher, and its relation to.
  43. [43]
    Disquisitiones Arithmeticae/Third Section - Wikisource
    Mar 28, 2024 · 55, II) that if p − 1 {\textstyle p-1} {\textstyle p-1} is = a α b ... {\textstyle p,} and therefore divides p − 1. {\textstyle p-1 ...
  44. [44]
    Augustin-Louis Cauchy - Biography - University of St Andrews
    Augustin-Louis Cauchy pioneered the study of analysis, both real and complex, and the theory of permutation groups. He also researched in convergence and ...Missing: subgroups | Show results with:subgroups
  45. [45]
    [PDF] A Course in Finite Group Representation Theory
    This course covers character theory of finite groups, representations over rings, and is for students beyond a first abstract algebra course, including topics ...
  46. [46]
    [PDF] Representation Theory - Berkeley Math
    ... degree of an irreducible representation divides the order of the group. We ... If G has order paqb, there exists an irreducible character χ and an element.Missing: source | Show results with:source<|control11|><|separator|>
  47. [47]
    [1105.5572] Lagrange's Theorem for Hopf Monoids in Species - arXiv
    May 27, 2011 · Abstract:Following Radford's proof of Lagrange's theorem for pointed Hopf algebras, we prove Lagrange's theorem for Hopf monoids in the category ...Missing: generalization | Show results with:generalization