Fact-checked by Grok 2 weeks ago

Subgroup

In group theory, a subgroup of a group G is a nonempty H of G that forms a group under the same as G. To qualify as a subgroup, H must satisfy under the (if a, b \in H, then ab \in H), contain the of G, and include the inverse of every element in H (if a \in H, then a^{-1} \in H); associativity is inherited from G. The notation H \leq G indicates that H is a subgroup of G. Subgroups provide essential insights into the of groups by allowing the study of smaller groups within larger ones, facilitating classifications and decompositions. Trivial subgroups include the singleton containing only the \{e\} and the entire group G itself, while proper subgroups exclude these cases. The of any collection of subgroups of G is itself a subgroup, and the smallest subgroup containing a given X \subseteq G, denoted \langle X \rangle, is generated by finite products of elements from X and their inverses. Key types of subgroups include cyclic subgroups, generated by a single element and isomorphic to either the infinite or a finite of order dividing the generator's order; normal subgroups, which are invariant under conjugation by elements of G (i.e., gHg^{-1} = H for all g \in G) and enable the formation of groups; and Sylow p-subgroups, which are maximal subgroups of order a power of a prime p and are central to the for classifying finite groups. Subgroups underpin fundamental results such as , which states that the order of a subgroup divides the order of the group, and play crucial roles in homomorphisms, group extensions, , and applications to and .

Definition and Fundamentals

Definition of a Subgroup

In group theory, a subgroup of a group G is a non-empty H \subseteq G that is closed under the group operation and under taking inverses. Formally, for all a, b \in H, the product a \cdot b \in H, and for all a \in H, the inverse a^{-1} \in H. The requirement that H is non-empty ensures the presence of the identity element e of G. To see this, let h \in H; since H is closed under inverses, h^{-1} \in H, and thus h \cdot h^{-1} = e \in H. With the identity in H and closure under the operation, H inherits the group structure from G. Subgroups are classified as proper if H \neq G. The trivial subgroup \{e\} contains only the , and G itself is always a subgroup, called the improper subgroup.

Notation and Conventions

In group theory, the standard notation for indicating that H is a subgroup of a group G is H \leq G. This symbol emphasizes the algebraic structure beyond mere set inclusion, as H \subseteq G alone does not guarantee the subgroup properties. For the subgroup generated by a subset S of G, the common notation is \langle S \rangle or hSi, representing the smallest subgroup containing S. When dealing with group orders, the cardinality |H| denotes the number of elements in a finite subgroup H, while infinite groups lack this finite measure and are not assigned an order in the same way. Groups are typically denoted as (G, *), where * is the binary operation, though additive notation may be used for abelian groups like (\mathbb{Z}, +). Not every subset of a group qualifies as a subgroup, as the subset must satisfy closure, identity inclusion, and inverses under the group operation. For instance, the even integers form a subgroup of (\mathbb{Z}, +) but do not constitute a group under multiplication, since the multiplicative identity 1 is odd and not included. Special cases include the trivial subgroup \{e\}, consisting solely of the identity element e, and the improper subgroup G itself, which is always a subgroup but excluded from considerations of proper subgroups.

Subgroup Criteria

One-Step Subgroup Test

The one-step subgroup test provides a streamlined for verifying whether a nonempty H of a group G forms a subgroup. Specifically, H is a subgroup of G if and only if a^{-1}b \in H for all a, b \in H. This condition combines aspects of and inverses into a single verification step, leveraging the group's existing structure. To see why this implies the standard subgroup properties, first note that the operation in [H](/page/H+) inherits associativity from [G](/page/G). Setting a = b yields a^{-1}a = e \in [H](/page/H+), confirming the identity is in [H](/page/H+). For inverses, fix a \in [H](/page/H+) and set b = e, so a^{-1}e = a^{-1} \in [H](/page/H+). For closure under the group operation, take a, b \in [H](/page/H+); since b^{-1} \in [H](/page/H+), it follows that a b = a (b^{-1})^{-1} \in [H](/page/H+) by the test applied to a and b^{-1}. Thus, all subgroup axioms hold. This test is particularly efficient for subsets where computing or verifying a^{-1}b is straightforward, such as in finite groups where elements can be enumerated or in settings with natural inversion operations like matrix groups. It applies equally to infinite groups but may be less practical there if the subset lacks a simple description under the inverse-difference operation. A representative example is the set of rotations in the dihedral group D_n (for n \geq 3), which consists of elements \{ e, r, r^2, \dots, r^{n-1} \}, where r is rotation by $2\pi/n. To verify using the one-step test, take a = r^i and b = r^j in this set; then a^{-1}b = r^{-i} r^j = r^{j-i}, which is also a power of r between 0 and n-1, hence in the set. Thus, the rotations form a subgroup isomorphic to the cyclic group of order n.

Two-Step Subgroup Test

The two-step subgroup test is a for verifying that a nonempty H of a group G forms a subgroup by separately confirming under the group operation and under inverses. Specifically, H \leq G if for all a, b \in H, ab \in H, and for all a \in H, a^{-1} \in H. This approach contrasts with more unified tests by emphasizing distinct verifications, which aids in building intuitive understanding of subgroup properties. The presence of the identity element e of G in H follows directly from these conditions. Since H is nonempty, select any h \in H; then h^{-1} \in H by inverse closure, and h \cdot h^{-1} = e \in H by operational closure. Associativity inherits from G, completing the structure. This test proves effective for manual verification in small or finite sets, where checking each property individually clarifies failures and successes without requiring advanced machinery. A counterexample illustrating the need for inverse closure is the set of positive real numbers in the additive group (\mathbb{R}, +). This set is nonempty and closed under addition, as the sum of positives remains positive, but it lacks inverses, since the additive inverse of any positive real is negative and outside the set.

Core Properties

Closure and Identity in Subgroups

A subgroup H of a group G is defined such that it is closed under the group , meaning that for all h_1, h_2 \in H, the product h_1 \cdot h_2 also belongs to H. This property ensures that H remains invariant under repeated applications of the from G, forming a self-contained within G. For instance, if h \in H, then the positive powers h^n for n \in \mathbb{N} are obtained by successive multiplications and thus lie in H, highlighting how propagates the iteratively. The e_G of the parent group G serves as the e_H for the subgroup H, as it satisfies e_G \cdot h = h \cdot e_G = h for all h \in H, inheriting this role directly from G's structure. This shared underscores the of H within G, requiring no separate verification since any subgroup must include e_G to maintain group axioms. Consequently, powers of elements extend to all integers n \in \mathbb{Z}, with h^n \in H for h \in H, as negative exponents incorporate the operation alongside . Associativity in H is automatically inherited from G, as the operation on H is merely the restriction of G's associative binary operation, eliminating the need for separate checks. This inheritance preserves the algebraic consistency essential to . Regarding finite generation, closure facilitates the construction of subgroups generated by a subset X \subseteq G, denoted \langle X \rangle, which comprises all finite products derived from elements of X and their inverses under the operation, forming the smallest subgroup containing X. For a single element x \in G, the cyclic subgroup \langle x \rangle = \{ x^n \mid n \in \mathbb{Z} \} exemplifies this, built solely through and the .

Inverses and Order Preservation

A fundamental property of subgroups is their under the : if [H](/page/H+) is a of a group G and h \in [H](/page/H+), then the h^{-1} must also belong to [H](/page/H+). This ensures that [H](/page/H+) forms a group under the same as G, allowing every in [H](/page/H+) to pair with its within [H](/page/H+) to yield the . Furthermore, the operation in subgroups inherits the general group property that the of an recovers the original : if h \in H, then (h^{-1})^{-1} = h. To see this, note that h^{-1} \cdot h = e and h \cdot h^{-1} = e by the group axioms; thus, h acts as a left and right for h^{-1}. Since inverses are in any group, it follows that (h^{-1})^{-1} = h, and both h and h^{-1} remain in H. This bidirectional reinforces the self-contained nature of subgroups. Subgroups also preserve the orders of their elements relative to the ambient group. Specifically, if a \in H \subseteq G and the of a in G is n (the smallest positive such that a^n = e), then the of a in H is also n, as the cyclic subgroup \langle a \rangle = \{ e, a, a^2, \dots, a^{n-1} \} is contained in H due to under the operation. More precisely, since the powers of a up to n generate the relations defining the , and all such powers lie in H, no smaller exponent yields the in H than in G. This equality holds because the minimal exponent is determined by the element's action alone, independent of the larger group structure. In finite groups, this order preservation has structural implications: the orders of elements in H must divide the order of G, teasing the deeper result that |H| divides |G|, though the full proof involves cosets. For instance, consider the rotation subgroup of the D_5, which models symmetries of a regular pentagon and consists of rotations R_0, R_1, R_2, R_3, R_4 (adding 0 through 4 modulo 5). The inverse of R_1 is R_4, since R_1 \circ R_4 = R_5 = R_0 (the ), and both remain in the rotation subgroup; similarly, R_2^{-1} = R_3. This illustrates inverse closure while preserving the of 5 for generators like R_1.

Cosets and Structural Theorems

Left and Right Cosets

In group theory, given a group G and a subgroup H \leq G, the left coset of H generated by an element a \in G is the set aH = \{ ah \mid h \in H \}. Similarly, the right coset is Ha = \{ ha \mid h \in H \}. These sets represent translates of H under left or right multiplication by a, respectively, and each has the same cardinality as H. Two left cosets aH and bH are equal if and only if a^{-1}b \in H. The analogous condition holds for right cosets: Ha = Hb if and only if ab^{-1} \in H. Moreover, distinct cosets are disjoint; that is, if aH \neq bH, then aH \cap bH = \emptyset. The collection of all left cosets of H in G forms a of G, meaning G is the of these cosets, each of size |H|. The same holds for right cosets. If G is finite, the number of distinct left (or right) cosets, denoted [G : H], equals |G| / |H|. A subgroup H is normal in G if and only if every left of H equals the corresponding right coset, i.e., aH = Ha for all a \in G. This equivalence holds automatically in abelian groups but requires additional structure otherwise.

Lagrange's Theorem and Consequences

Lagrange's theorem is a fundamental result in group theory that relates the orders of a finite group and its subgroups. It states that if G is a finite group and H is a subgroup of G, then the order of H, denoted |H|, divides the order of G, denoted |G|. Additionally, the index of H in G, written [G : H], which is the number of distinct left (or right) cosets of H in G, equals |G| / |H|. The proof relies on the of G into disjoint left of H. Each gH = \{ gh \mid h \in H \} for g \in G has exactly |H| elements, as the map h \mapsto gh is a by left cancellation in groups. Distinct cosets are disjoint, and their union is G, so |G| equals the number of cosets times |H|, implying |H| divides |G| and [G : H] = |G| / |H|. A key corollary is that the order of any element g \in G divides |G|, since the cyclic subgroup \langle g \rangle generated by g has order equal to the order of g, which must divide |G| by Lagrange's theorem. Consequently, g^{|G|} = e for all g \in G, where e is the identity. Another important consequence is that if H is a proper subgroup of another subgroup K \leq G, then |K| \geq 2 |H|, because K is a disjoint union of at least two cosets of H in K, so no subgroup of G can have order strictly between |H| and $2 |H|. Lagrange's theorem has significant applications in . For instance, —that if p is prime and a is an not divisible by p, then a^{p-1} \equiv 1 \pmod{p}—follows as a special case. The (\mathbb{Z}/p\mathbb{Z})^\times of units p has order p-1, so by the on element orders, the order of _p divides p-1, yielding _p^{p-1} = {{grok:render&&&type=render_inline_citation&&&citation_id=1&&&citation_type=wikipedia}}_p. While powerful for finite groups, Lagrange's theorem does not extend to infinite groups in the same way, as subgroup orders need not "divide" the group's cardinality in a constraining manner. For example, the additive group \mathbb{Z} of integers has subgroups m\mathbb{Z} for each positive integer m, all of which are infinite and have finite index m, but the theorem's divisibility condition on finite orders does not apply or impose similar restrictions.

Illustrative Examples

Subgroups of the ℤ₈

The \mathbb{Z}_8 is the additive group of integers 8, consisting of the elements \{0, 1, 2, 3, 4, 5, 6, 7\} with 8. As a finite of order 8, its subgroups are precisely the cyclic subgroups generated by elements whose orders divide 8, and there is exactly one subgroup for each divisor of 8. The trivial subgroups are the identity subgroup \{0\} of order 1 and the full group \mathbb{Z}_8 itself of order 8. The proper nontrivial subgroups are \langle 4 \rangle = \{0, 4\} of order 2 and \langle 2 \rangle = \{0, 2, 4, 6\} of order 4; note that \langle 1 \rangle = \mathbb{Z}_8, \langle 3 \rangle = \mathbb{Z}_8, \langle 5 \rangle = \mathbb{Z}_8, \langle 7 \rangle = \mathbb{Z}_8, \langle 6 \rangle = \langle 2 \rangle, and \langle 0 \rangle = \{0\}. There are no subgroups of order 3 or 5, as the possible orders of subgroups must divide 8 by . All subgroups of \mathbb{Z}_8 are cyclic, as are always cyclic, and they are generated by multiples of the generator corresponding to the s of 8 (specifically, generated by k where k = 8/d for each d of 8). The lattice of \mathbb{Z}_8 forms a chain reflecting the lattice of 8: \{0\} \subset \langle 4 \rangle \subset \langle 2 \rangle \subset \mathbb{Z}_8, where each is proper and the structure arises from the unique ordered by corresponding to the 1, 2, 4, and 8.
SubgroupGeneratorOrderElements
Trivial\langle [0](/page/0) \rangle or \langle 8 \rangle\{[0](/page/0)\}
Order 2\langle 4 \rangle2\{0, 4\}
Order 4\langle 2 \rangle4\{0, 2, 4, 6\}
Full group\langle [1](/page/1) \rangle8\{0, 1, 2, 3, 4, 5, 6, 7\}

Subgroups of the Symmetric Group S₄

The symmetric group S_4, consisting of all permutations of four elements, has order $4! = 24. By , the possible orders of its subgroups are the divisors of 24: 1, 2, 3, 4, 6, 8, 12, and 24. This classification highlights the rich subgroup structure of S_4, including both abelian and non-abelian examples, with multiple types for some orders. The subgroups can be enumerated up to conjugacy, revealing distinct conjugacy classes based on types and . The trivial subgroup of order 1 is unique, consisting solely of the identity permutation. Subgroups of order 2 are cyclic, generated by elements of order 2; there are two es: six subgroups generated by transpositions (cycle type 2,1,1), such as \langle (1\,2) \rangle = \{ e, (1\,2) \}, and three generated by double transpositions (cycle type 2,2), such as \langle (1\,2)(3\,4) \rangle = \{ e, (1\,2)(3\,4) \}. For order 3, all subgroups are cyclic, generated by 3-cycles; there is one conjugacy class containing four such subgroups, exemplified by \langle (1\,2\,3) \rangle = \{ e, (1\,2\,3), (1\,3\,2) \}. Order 4 subgroups fall into three types up to conjugacy: three cyclic subgroups generated by 4-cycles, such as \langle (1\,2\,3\,4) \rangle = \{ e, (1\,2\,3\,4), (1\,3)(2\,4), (1\,4\,3\,2) \}; one transitive V_4 = \{ e, (1\,2)(3\,4), (1\,3)(2\,4), (1\,4)(2\,3) \}, which is in S_4; and three intransitive s, such as \langle (1\,2), (3\,4) \rangle = \{ e, (1\,2), (3\,4), (1\,2)(3\,4) \}. Subgroups of order 6 are isomorphic to S_3; there is one containing four such subgroups, each the of a point, acting as the full on the remaining three elements, for example, the of 4 is \{ e, (1\,2), (1\,3), (2\,3), (1\,2\,3), (1\,3\,2) \}. For order 8, the Sylow 2-subgroups are isomorphic to the D_4 of order 8 (symmetries of the square), and there are three conjugate copies, such as the one generated by rotations and reflections in the permutation , exemplified by \{ e, (1\,3\,2\,4), (1\,4\,2\,3), (1\,2)(3\,4), (1\,4)(2\,3), (1\,3)(2\,4), (1\,2), (3\,4) \}; their number follows from Sylow's theorems, as the equals the number of Sylow 2-subgroups, [S_4 : N_{S_4}(P)] = 3 where P is a Sylow 2-subgroup. The unique subgroup of order 12 is the A_4, consisting of all even , such as A_4 = \{ e \} \cup \{ all 3-cycles and double transpositions \}; its uniqueness follows from the fact that any order-12 subgroup must contain all elements of order 3 (the eight 3-cycles) and then all double transpositions to close under multiplication. Finally, the whole group S_4 is the unique subgroup of order 24.

Additional Examples and Applications

Subgroups in Additive Groups of Integers

The additive group of , denoted (\mathbb{Z}, +), is an infinite generated by 1. Its subgroups play a fundamental role in understanding the structure of abelian groups and ideal theory in . Every subgroup H of \mathbb{Z} under addition is of the form n\mathbb{Z} = \{nk \mid k \in \mathbb{Z}\} for some nonnegative integer n \geq 0. The trivial subgroup is $0\mathbb{Z} = \{0\}, which consists solely of the identity element. The full group corresponds to $1\mathbb{Z} = \mathbb{Z}. For n > 1, n\mathbb{Z} comprises all multiples of n, forming a proper subgroup. These subgroups arise because \mathbb{Z} is a (PID), where every ideal—equivalently, every additive subgroup—is principal, generated by a single element n. In this context, the subgroup generated by a finite set of integers a_1, \dots, a_k is d\mathbb{Z}, where d = \gcd(a_1, \dots, a_k). All subgroups of \mathbb{[Z](/page/Z)} are infinite cyclic groups of infinite , as \mathbb{Z} is torsion-free: no nontrivial element has finite order. Consequently, there are no finite nontrivial subgroups. The [\mathbb{Z} : n\mathbb{Z}] equals n for n > 0, reflecting the n distinct cosets $0 + n\mathbb{Z}, 1 + n\mathbb{Z}, \dots, (n-1) + n\mathbb{Z}. A representative example is $2\mathbb{Z}, the subgroup of even integers \{\dots, -4, -2, 0, 2, 4, \dots\}. The of subgroups m\mathbb{Z} \cap n\mathbb{Z} = \operatorname{lcm}(m,n) \mathbb{Z}; for instance, $2\mathbb{Z} \cap 3\mathbb{Z} = 6\mathbb{Z}, the multiples of 6.

Subgroups in Matrix Groups

The general linear group \mathrm{GL}(n, \mathbb{R}) consists of all invertible n \times n real matrices under matrix multiplication, forming an open subgroup of the space of all n \times n matrices. Within this group, notable subgroups arise from geometric constraints or determinant conditions, illustrating how linear transformations preserve specific structures. For instance, the orthogonal group \mathrm{O}(n) is the subgroup of matrices A satisfying A^T A = I, which preserves the Euclidean norm \|v\| for all vectors v \in \mathbb{R}^n, as \|Av\| = \|v\|. Similarly, the special linear group \mathrm{SL}(n, \mathbb{R}) comprises matrices in \mathrm{GL}(n, \mathbb{R}) with determinant 1, preserving oriented volumes in \mathbb{R}^n. Another important class involves unipotent subgroups, such as the set of upper triangular matrices with 1s on the diagonal, denoted U(n), which forms a subgroup of \mathrm{GL}(n, \mathbb{R}). These matrices satisfy (I + N)^k = I + kN + \cdots where N is strictly upper triangular and , enabling applications in solving systems of linear differential equations via the . Cyclic subgroups also appear prominently; for example, in \mathrm{GL}(2, \mathbb{R}), the subgroup generated by a R_\theta = \begin{pmatrix} \cos \theta & -\sin \theta \\ \sin \theta & \cos \theta \end{pmatrix} with \theta = 2\pi / m for m yields a finite of order m, representing discrete rotations in the plane. If m is finite, this subgroup embeds into the special orthogonal group \mathrm{SO}(2), a of \mathrm{O}(2). The Borel subgroup B(n) consists of all invertible upper triangular matrices in \mathrm{GL}(n, \mathbb{R}), a maximal solvable subgroup that plays a central role in the Bruhat decomposition of \mathrm{GL}(n, \mathbb{R}). It decomposes as B(n) = T(n) \ltimes U(n), where T(n) is the diagonal matrices (torus) and U(n) the unipotent radical, facilitating the study of parabolic inductions in representation theory. For a finite example, consider \mathrm{GL}(2, \mathbb{Z}), the group of $2 \times 2 integer matrices with determinant \pm 1; its subgroup \mathrm{SL}(2, \mathbb{Z}) has index 2, and the modular group \mathrm{PSL}(2, \mathbb{Z}) = \mathrm{SL}(2, \mathbb{Z}) / \{\pm I\} is a discrete subgroup acting on the hyperbolic plane, generating fundamental domains for modular forms. In the context of Lie groups, \mathrm{GL}(n, \mathbb{R}) itself is a Lie group, and its closed subgroups like \mathrm{O}(n) and \mathrm{SL}(n, \mathbb{R}) inherit smooth manifold structures, while discrete subgroups such as \mathrm{SL}(2, \mathbb{Z}) provide arithmetic examples embedded in the continuous framework, essential for understanding rigidity and dynamics in higher-rank semisimple Lie groups. These subgroups highlight algebraic applications in geometry, such as classifying orbits under group actions, and in number theory, where discrete matrix subgroups model automorphic forms.

References

  1. [1]
    4.1: Introduction to Subgroups - Mathematics LibreTexts
    Sep 25, 2021 · Definition: Subgroup. A subgroup of a group G is a subset of G that is also a group under G 's operation. If H is a subgroup of G , we write ...
  2. [2]
    [PDF] Group Theory - James Milne
    Group theory is the study of symmetries. Definitions and examples. Definition 1.1 A group is a set 𝐺 together with a mapping. (𝑎, 𝑏) ↦ 𝑎 ∗ 𝑏∶ 𝐺 × 𝐺 ...
  3. [3]
    Évariste Galois (1811 - 1832) - Biography - MacTutor
    Évariste Galois was a French mathematician who produced a method of determining when a general equation could be solved by radicals and is famous for his ...
  4. [4]
    [PDF] Algebra I - Mathematisches Seminar
    If (G, ◦) is a group, we write H ≤ G to say that H is a subgroup of (G, ◦) if due to the context there is no doubt about the operation ◦ to which this statement ...
  5. [5]
    [PDF] Abstract Algebra I - Auburn University
    Nov 15, 2018 · the diagram again using just the cyclic subgroup notation: Z12 h2i h3i h4i h6i. {0} ... (i) If H ≤ G, then ϕ(H) ≤ G0. (ii) If H0 ≤ G0, then ϕ−1(H0) ...
  6. [6]
    [PDF] subgroups.pdf
    Jan 16, 2018 · It's common to write hSi for the subgroup generated by S. So in case S = {x1,x2,...,xn} (a finite set), write hx1,x2,.
  7. [7]
    [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 ...
  8. [8]
    [PDF] Section I.5. Subgroups
    Jul 7, 2023 · Just as a vector space can have a subspace, as you see in linear algebra, a group can have a subgroup. As with vector spaces, this would be a ...
  9. [9]
    [PDF] Math 403 Chapter 3: Finite Groups and Subgroups
    Example: Z5 is not a subgroup of Z. It is a subset but the operations are different. 3. Special Subgroups: There are certain subgroups of groups which will be ...
  10. [10]
    [PDF] dihedral groups - keith conrad
    We will look at elementary aspects of dihedral groups: listing its elements, relations between rotations and reflections, the center, and conjugacy classes.
  11. [11]
  12. [12]
    [PDF] 18.703 Modern Algebra, Subgroups - MIT OpenCourseWare
    Then H is closed under addition (the sum of two positive numbers is positive) but the inverse of every element of H does not lie in H. Definition 2.2. Let G ...
  13. [13]
    [PDF] Abstract Algebra
    a divides b the greatest common divisor of a, b also the ideal generated by a, b the order of the set A, the order of the element x.
  14. [14]
    [PDF] 4 Proofs in group theory
    (a) understand that the identity in a group is unique;. (b) understand that each element in a group has a unique inverse;. (c) recognise how the uniqueness ...
  15. [15]
    [PDF] ORDERS OF ELEMENTS IN A GROUP 1. Introduction Let G be a ...
    An element g in a group has finite order if gn=e for some positive integer n. The order of g is the least n such that gn=e. If no such n exists, g has infinite ...
  16. [16]
    [PDF] 16 Cyclic and Dihedral Groups
    To verify that Dn is closed under inverses, first note that every rotation Ra has inverse R-a. We claim that every mirror Ma is its own inverse. To check ...
  17. [17]
    6.1: Cosets - Mathematics LibreTexts
    Jun 4, 2022 · In a commutative group, left and right cosets are always identical. Example ...
  18. [18]
    5.1: Cosets - Mathematics LibreTexts
    Apr 16, 2022 · Definition: Left and Right Cosets. Let G be a group and let H ≤ G and a ∈ G . The subsets (5.1.1) a ⁢ H := { a ⁢ h ∣ h ∈ H } and ...Missing: theory | Show results with:theory
  19. [19]
    [PDF] Lagrange's Theorem: Statement and Proof - St. Olaf College
    Apr 5, 2002 · If G is a group with subgroup H, then there is a one to one correspondence between H and any coset of H. Proof. Let C be a left coset of H in G.Missing: corollaries | Show results with:corollaries
  20. [20]
    [PDF] 10 Cosets and Lagrange's Theorem - UC Berkeley math
    Lagrange's Theorem is one of the most important combinatorial results in finite group theory and will be used repeatedly. Corollary. If G is a group of prime ...
  21. [21]
    [PDF] Cosets and Lagrange's theorem - Keith Conrad
    Corollary 7.9. Every group of prime size is cyclic, and in fact each non-identity element is a generator. Proof. Let G be the ...Missing: statement | Show results with:statement
  22. [22]
    [PDF] MATH 433 Applied Algebra Lecture 29: Cosets. Lagrange's Theorem.
    Corollary 3 (Fermat's Little Theorem) If p is a prime number then ap-1 ≡ 1 mod p for any integer a that is not a multiple of p. Proof: ap-1 ≡ 1 mod p means that ...
  23. [23]
    [PDF] cyclic-groups.pdf
    ... 8, 12, and 24, the subgroups of Z24 are: h0i, h1i, h2i, h3i, h4i, h6i, h8i, h12i. The subgroup generated by 3 has order 8: h3i = {0, 3, 6, 9, 12, 15, 18, 21} ...
  24. [24]
  25. [25]
    subgroups of S_4 - PlanetMath
    Mar 22, 2013 · But any such element together with a 3 3 -cycle generates A4 A 4 . Thus A4 A 4 is the only subgroup of S4 S 4 of order 12 12 .
  26. [26]
    [PDF] Subgroups of cyclic groups - Keith Conrad
    By Theorem 2.1, every subgroup of Z (additive!) has the form nZ for some integer n, and we can take n ≥ 0. Obviously different choices of n ≥ 0 give us ...
  27. [27]
    IAAWA Principal Ideal Domains - UTK Math
    An integral domain in which every ideal is principal is called a principal ideal domain , or PID . 🔗. Theorem 7.19.<|control11|><|separator|>
  28. [28]
    [PDF] Problem 1. Let I = nZ and J = mZ be ideals of Z. - People
    Let us apply this to the case when R = Z, I = nZ, J = mZ for some ralatively prime integers m, n. By a) of the previous problem, we have I + J = gcd(m, n)Z = Z ...
  29. [29]
    [PDF] Chapter 3 Review of Groups and Group Actions - UPenn CIS
    This group is called the general linear group and is usually denoted by GL(n,R) (or GL(n,C)). 9. The set of n×n invertible matrices with real (or com- plex) ...
  30. [30]
    [PDF] GLn(R) AS A LIE GROUP Contents 1. Matrix Groups over R, C, and ...
    matrices of positive determinant is called the special orthogonal group. Note that it is equivalent to define the orthogonal group as the subgroup of. GLn(R) ...<|control11|><|separator|>
  31. [31]
    [PDF] LINEAR LIE GROUPS 1. The general linear group Recall that M(n,R ...
    By definition they are Lie subgroups of GL(n,R). Example. (The special linear group) The special linear group is defined as. SL(n,R) = {X ∈ GL(n,R) : detX = 1}.
  32. [32]
    [PDF] PRIME POWERS UNITS AND FINITE SUBGROUPS OF GLn(Q)
    For example, if d if a positive integer then a counterclockwise rotation by 2π/d radians in the plane R2 is represented by the matrix cos(2π/d) −sin(2π/d) sin( ...
  33. [33]
    [PDF] sl2(z) - keith conrad
    Proof of Theorem 1.1. Let G = hS, Ti be the subgroup of SL2(Z) generated by S and T. We will give two proofs that G = SL2(Z), one algebraic and the other ...
  34. [34]
    Discrete subgroups of Lie groups and discrete transformation groups
    Aug 10, 2011 · They provide special, important examples of lattices of semisimple Lie groups and also nonlattices. Crystallographic groups (Bieberbach theorems) ...
  35. [35]
    [PDF] GENERAL PARABOLIC SUBGROUPS OF GLn(C)
    We wish to show that all Borel subgroups of GLn(C) are conjugate to the subgroup B of invertible n×n upper triangular matrices over C. To verify this we ...