Fact-checked by Grok 2 weeks ago

Group isomorphism

In , a group isomorphism is a bijective between two groups, meaning it is a and onto that preserves the group operation by mapping the product of any two elements in the first group to the product of their images in the second group. Groups related by an isomorphism are denoted as isomorphic (often written G \cong H) and possess identical algebraic structures, differing only in the notation or labeling of their elements. Isomorphisms form an equivalence relation on the class of all groups, being reflexive (every group is isomorphic to itself via the identity map), symmetric (if G \cong H, then H \cong G via the inverse map), and transitive (if G \cong H and H \cong K, then G \cong K via the composition of maps). They preserve fundamental properties, including the order of the group, the orders of individual elements, the , inverses, abelianness, cyclicity, and the existence and orders of subgroups. This preservation ensures that isomorphic groups are indistinguishable in terms of their abstract behavior and structural features. A central goal in group theory is the classification of groups up to isomorphism, which seeks to identify and enumerate all distinct group structures by considering isomorphic groups as equivalent. Notable examples include the isomorphism between the additive group of real numbers (\mathbb{R}, +) and the multiplicative group of positive real numbers (\mathbb{R}^+, \times) via the exponential function f(x) = e^x, demonstrating how different operations can underlie the same structure. For small finite orders, such as 1, 2, or 3, there is a unique group up to isomorphism in each case, highlighting the finiteness of isomorphism classes for low orders.

Core Definitions

Group Homomorphisms

A is a \phi: G \to H between two groups G and H that preserves the group operation, meaning \phi(ab) = \phi(a)\phi(b) for all a, b \in G./11:_Homomorphisms/11.01:_Group_Homomorphisms) This condition implies that \phi maps the identity element e_G of G to the e_H of H, since setting a = b = e_G yields \phi(e_G) = \phi(e_G)^2, so \phi(e_G) must be e_H. Homomorphisms preserve additional structural features of groups. Specifically, if an element a \in G has finite order n, then the order of \phi(a) divides n, because if a^n = e_G, applying \phi gives \phi(a)^n = e_H./11:_Homomorphisms/11.01:_Group_Homomorphisms) Moreover, the image of any subgroup K \leq G under \phi, denoted \phi(K) = \{\phi(k) \mid k \in K\}, forms a subgroup of H. The kernel of a homomorphism \phi: G \to H, denoted \ker \phi = \{g \in G \mid \phi(g) = e_H\}, is the set of elements in G that map to the identity in H. This kernel is always a normal subgroup of G, as for any g \in G and k \in \ker \phi, one has \phi(gkg^{-1}) = \phi(g)\phi(k)\phi(g)^{-1} = \phi(g)e_H\phi(g)^{-1} = e_H, so gkg^{-1} \in \ker \phi./11:_Homomorphisms/11.01:_Group_Homomorphisms) The image of \phi, denoted \operatorname{im} \phi = \{\phi(g) \mid g \in G\}, is a subgroup of H. For finite groups, the order of G satisfies |G| = |\ker \phi| \cdot |\operatorname{im} \phi|, which follows from the fact that the cosets of \ker \phi in G are in bijection with the elements of \operatorname{im} \phi, applying Lagrange's theorem to the index of the kernel. Isomorphisms are precisely the bijective group homomorphisms./11:_Homomorphisms/11.01:_Group_Homomorphisms)

Group Isomorphisms

A group isomorphism is a bijective group homomorphism \phi: G \to H between two groups G and H. While homomorphisms preserve the group operation in one direction, isomorphisms ensure a two-way structural correspondence due to their bijectivity. For such a \phi to qualify as an isomorphism, its inverse \phi^{-1}: H \to G must also be a homomorphism. To see this, let a', b' \in H; set a = \phi^{-1}(a') and b = \phi^{-1}(b'). Then \phi^{-1}(a'b') = \phi^{-1}(\phi(a)\phi(b)) = \phi^{-1}(\phi(ab)) = ab = \phi^{-1}(a')\phi^{-1}(b'), confirming that \phi^{-1} preserves the operation. Additionally, since \phi maps the identity e_G to e_H, the inverse maps e_H back to e_G. If an isomorphism exists between G and H, the groups are denoted G \cong H and are said to be structurally equivalent, meaning they share the same abstract group structure despite potentially different underlying sets. This equivalence implies that G and H cannot be distinguished by any intrinsic group-theoretic properties. Isomorphisms preserve all fundamental group-theoretic properties. Specifically, they preserve the order of the group |G| = |H| due to bijectivity, and the order of elements |g| = |\phi(g)| for all g \in G, as powers and the identity are mapped accordingly. Subgroups of G correspond bijectively to subgroups of H via \phi, since the image of a subgroup under a homomorphism is a subgroup, and bijectivity ensures a one-to-one matching. Similarly, cosets are preserved: the cosets of a subgroup K \leq G map to the cosets of \phi(K) \leq H. Relations between elements, such as commutators [g, h] = g^{-1}h^{-1}gh mapping to [\phi(g), \phi(h)], are also maintained, ensuring that properties like abelianness hold equivalently in both groups.

Fundamental Properties

Preservation of Structure

An isomorphism \phi: G \to H between groups G and H preserves the group operation pointwise, meaning \phi(ab) = \phi(a)\phi(b) for all a, b \in G. This property directly implies that \phi maps the identity element of G to the identity of H and inverses to inverses, ensuring the structural integrity of the operation is maintained across the groups. Extending this, powers are preserved such that \phi(a^n) = \phi(a)^n for any n, as the operation's associativity and the homomorphism property allow iterative application. Consequently, the order of an element is invariant: if a \in G has finite order m, then \phi(a) has order m in H. Beyond elements, isomorphisms preserve the subgroup lattice bijectively: if K is a of G, then \phi(K) is a of H, and the inverse map \phi^{-1} ensures the correspondence is . This bijectivity stems from the isomorphism's invertibility, mapping to without loss or addition of structure. Moreover, are preserved: if K \trianglelefteq G, then \phi(K) \trianglelefteq H, as the normality condition—conjugation by group elements—translates under the structure-preserving map. Several key invariants remain unchanged under isomorphism. The center Z(G) = \{ z \in G \mid zg = gz \ \forall g \in G \} of G maps to the center Z(H) of H, since it is a characteristic subgroup fully determined by the group's operation. Similarly, the derived subgroup G' = \langle [g,h] \mid g,h \in G \rangle, generated by commutators, is preserved as \phi(G') = H', reflecting its role as the smallest normal subgroup capturing non-abelian structure. The abelianization G/G', the largest abelian quotient of G, is thus isomorphic to H/H'. The exponent of the group, defined as the least common multiple of the orders of its elements (or infinite if unbounded), is also invariant, following from the preservation of individual element orders. For cyclic subgroups, an isomorphism satisfies \langle \phi(a) \rangle = \phi(\langle a \rangle) for any a \in G, with generators mapping to generators: if a generates \langle a \rangle, then \phi(a) generates \phi(\langle a \rangle). This ensures that the cyclic structure, including relations among powers, is fully transferred.

Isomorphisms as Equivalence

In group theory, the relation of defines an on the class of all groups. Specifically, two groups G and H are related if there exists a group \phi: G \to H. This relation is reflexive, symmetric, and transitive. To see that the relation is reflexive, consider the identity map \mathrm{id}_G: G \to G defined by \mathrm{id}_G(g) = g for all g \in G. This map is a , and it preserves the group since \mathrm{id}_G(gg') = gg' = \mathrm{id}_G(g) \mathrm{id}_G(g') for all g, g' \in G. Thus, \mathrm{id}_G is an , so G \cong G. The relation is symmetric: if \phi: G \to H is an isomorphism, then \phi^{-1}: H \to G exists as a . Moreover, \phi^{-1} preserves the operation because \phi^{-1}(h h') = \phi^{-1}(h) \phi^{-1}(h') follows from applying \phi to both sides and using the homomorphism property of \phi. Hence, H \cong G. Transitivity holds as follows: if \phi: G \to H and \psi: H \to K are isomorphisms, then the \psi \circ \phi: G \to K is a (since both \phi and \psi are bijections). It preserves the because (\psi \circ \phi)(g g') = \psi(\phi(g g')) = \psi(\phi(g) \phi(g')) = \psi(\phi(g)) \psi(\phi(g')) = (\psi \circ \phi)(g) (\psi \circ \phi)(g') for all g, g' \in G. Thus, G \cong K. This partitions the class of all groups into equivalence classes, known as isomorphism classes. The isomorphism class of a group G, denoted [G], consists of all groups H such that H \cong G. Groups within the same isomorphism class are structurally identical, meaning they share all group-theoretic properties preserved by isomorphisms, such as , the existence of certain subgroups, or solvability. The concept of isomorphism classes has profound implications for the of groups. Classification efforts aim to describe groups up to using invariants, such as the group's , whether it is abelian or non-abelian, or its by generators and relations. Groups in the same class are considered indistinguishable, allowing mathematicians to focus on representatives of each class rather than enumerating all possible groups. This approach underpins major results, like the complete .

Illustrative Examples

Basic Examples

A fundamental example of a group isomorphism involves the additive group of integers and its subgroups. Consider the map \phi: (\mathbb{Z}, +) \to (n\mathbb{Z}, +) defined by \phi(x) = n x, where n is a fixed non-zero integer. This map is a homomorphism because \phi(x + y) = n (x + y) = n x + n y = \phi(x) + \phi(y). It is injective, as the kernel is trivial: if \phi(x) = 0, then n x = 0, so x = 0 since \mathbb{Z} has no zero divisors. It is also surjective, as every element n k in n\mathbb{Z} is \phi(k) for k \in \mathbb{Z}. Thus, \phi establishes that (\mathbb{Z}, +) \cong (n\mathbb{Z}, +). Another basic example is the isomorphism between the Klein four-group V_4 = \{e, a, b, ab\}, where a^2 = b^2 = e and ab = ba, and the direct product \mathbb{Z}_2 \times \mathbb{Z}_2 = \{(0,0), (1,0), (0,1), (1,1)\} with componentwise addition modulo 2. The explicit isomorphism is given by \phi(e) = (0,0), \phi(a) = (1,0), \phi(b) = (0,1), and \phi(ab) = (1,1). To verify it is a homomorphism, note that, for instance, \phi(a \cdot a) = \phi(e) = (0,0) and \phi(a) + \phi(a) = (1,0) + (1,0) = (0,0); similarly for other products like a \cdot b = ab mapping to (1,0) + (0,1) = (1,1). The map is bijective, as it pairs each element uniquely and covers all four elements of the codomain, with trivial kernel confirming injectivity. For finite cyclic groups, the additive group \mathbb{Z}_4 = \{0, 1, 2, 3\} with addition modulo 4 is isomorphic to itself via the identity map \phi(k) = k, which trivially preserves the operation, is injective and surjective, and has 4 like 1 with 4. In contrast, \mathbb{Z}_4 is not isomorphic to \mathbb{Z}_2 \times \mathbb{Z}_2, as the latter has all non-identity of 2 (e.g., (1,0) + (1,0) = (0,0)), while \mathbb{Z}_4 has an of 4, violating preservation of orders under any potential .

Non-Trivial Examples

One prominent non-trivial example of a group isomorphism is between the S_3, consisting of all permutations of three elements, and the D_3 of order 6, which describes the symmetries of an . The isomorphism maps the permutations in S_3 to compositions of rotations and reflections in D_3; specifically, the maps to the , the 3-cycles (123) and (132) map to rotations by 120° and 240° respectively, and the transpositions (12), (13), (23) map to reflections across the corresponding axes. This preserves the group operation, as verified by checking that the relations in both presentations hold equivalently. In the infinite case, the F_2 on two generators embeds as a of the SL(2, \mathbb{Z}), which consists of $2 \times 2 integer matrices with determinant 1. A realization is the Sanov generated by the matrices \begin{pmatrix} 1 & 2 \\ 0 & 1 \end{pmatrix} and \begin{pmatrix} 1 & 0 \\ 2 & 1 \end{pmatrix}, which is free of rank 2, establishing the F_2 \cong \langle A, B \rangle \leq SL(2, \mathbb{Z}). Another striking infinite example involves the additive group of real numbers (\mathbb{R}, +), which is to a proper of itself. Such an can be constructed non-constructively using Hamel bases over \mathbb{Q}: both \mathbb{R} and the proper have Hamel bases of ; extend a between the bases linearly to obtain a \mathbb{Q}-linear . However, this relies on the to guarantee the existence of such bases. A non-abelian example highlighting structural complexity is the embedding of the special orthogonal group SO(3), the group of 3D rotations preserving orientation, as a subgroup of the general linear group GL(3, \mathbb{R}). Explicitly, elements of SO(3) are represented by rotation matrices, such as the rotation by angle \theta around the z-axis given by \begin{pmatrix} \cos \theta & -\sin \theta & 0 \\ \sin \theta & \cos \theta & 0 \\ 0 & 0 & 1 \end{pmatrix}, which satisfy A^T A = I and \det A = 1, confirming the subgroup property under matrix multiplication. These examples appear non-trivial because the groups often admit vastly different presentations or geometric interpretations, yet they share identical multiplication tables (for finite cases) or structural properties up to relabeling of elements, underscoring that equates abstract structures despite concrete dissimilarities.

First

The First , also known as the Homomorphism , establishes a deep connection between group homomorphisms, their s, and quotient groups. It states that if \phi: G \to H is a , then the K = \ker \phi is a of G, and there is an \overline{\phi}: G/K \to \operatorname{im} \phi given by \overline{\phi}(gK) = \phi(g) for all g \in G. To verify that \overline{\phi} is well-defined, suppose g_1 K = g_2 K for some g_1, g_2 \in [G](/page/G). Then g_1^{-1} g_2 \in K, so \phi(g_1^{-1} g_2) = [e_H](/page/Identity), where e_H is the identity in H. Thus, \phi(g_1) = \phi(g_2), implying \overline{\phi}(g_1 K) = \overline{\phi}(g_2 K). Next, \overline{\phi} is a homomorphism because it preserves the group operation: for cosets g_1 K, g_2 K \in G/K, \overline{\phi}(g_1 K \cdot g_2 K) = \overline{\phi}(g_1 g_2 K) = \phi(g_1 g_2) = \phi(g_1) \phi(g_2) = \overline{\phi}(g_1 K) \cdot \overline{\phi}(g_2 K). This confirms that \overline{\phi} respects the multiplication in the quotient group and the image subgroup. The map \overline{\phi} is injective: if \overline{\phi}(g_1 K) = \overline{\phi}(g_2 K), then \phi(g_1) = \phi(g_2), so \phi(g_1^{-1} g_2) = e_H and g_1^{-1} g_2 \in K, hence g_1 K = g_2 K. It is surjective onto \operatorname{im} \phi because for any h \in \operatorname{im} \phi, there exists g \in G such that \phi(g) = h, so \overline{\phi}(g K) = h. Therefore, \overline{\phi} is an isomorphism. A key is that every homomorphic image of a group G is isomorphic to some of G. Specifically, \operatorname{im} \phi \cong G / \ker \phi, which provides a powerful method for classifying groups up to by identifying them with familiar structures.

Second and Third

The second isomorphism theorem provides a relationship between a subgroup and a normal subgroup of a group, describing an isomorphism between certain quotient groups. Let G be a group, H \leq G a subgroup, and N \trianglelefteq G a normal subgroup. Then HN = \{hn \mid h \in H, n \in N\} is a subgroup of G, N \trianglelefteq HN, and H \cap N \trianglelefteq H. Moreover, there is a group isomorphism HN / N \cong H / (H \cap N) given by the map \phi: H \to HN / N defined by \phi(h) = hN, which is well-defined because N \trianglelefteq G. To see that \phi is a homomorphism, note that for h_1, h_2 \in H, \phi(h_1 h_2) = h_1 h_2 N = (h_1 N)(h_2 N) = \phi(h_1) \phi(h_2), since coset multiplication is preserved. The map \phi is surjective because every element of HN / N is of the form h'n' for some h' \in H, n' \in N, which equals h'N = \phi(h'). The kernel of \phi is \ker \phi = \{h \in H \mid hN = N\} = H \cap N. By the first isomorphism theorem applied to \phi, the image is isomorphic to H / \ker \phi = H / (H \cap N), and since the image is all of HN / N, the desired isomorphism holds. This establishes the coset equality and normality preservation in the product subgroup. The third isomorphism theorem extends this by relating quotients in a chain of normal subgroups, showing compatibility under successive factoring. Let G be a group with normal subgroups N \trianglelefteq K \trianglelefteq G. Then K / N \trianglelefteq G / N, and there is a group (G / N) / (K / N) \cong G / K induced by the natural projection \pi: G \to G / K given by \pi(g) = gK, which factors through G / N. Specifically, the isomorphism \psi: (G / N) / (K / N) \to G / K is defined by \psi((gN)(K / N)) = gK for g \in G. The \psi arises from the composition of the natural surjection G / N \to G / K (with K / N) and the G \to G / N. To verify, first define \overline{\pi}: G / N \to G / K by \overline{\pi}(gN) = gK; this is a well-defined because if gN = g'N, then g^{-1}g' \in N \subseteq K, so gK = g'K. It is surjective since every gK equals \overline{\pi}(gN). The is \{\, gN \mid gK = K \,\} = K / N. Applying the first theorem to \overline{\pi} yields (G / N) / \ker \overline{\pi} \cong G / K, which is the stated . This confirms the of K / N in G / N via containment: for g \in G, n \in N, k \in K, (gN)(kN)(gN)^{-1} = (gkg^{-1})N \in K / N since K \trianglelefteq G and N \subseteq K.

Special Cases

Cyclic Groups

A cyclic group is generated by a single element, and all finite cyclic groups of the same order are isomorphic to each other and to the additive group \mathbb{Z}_n. Specifically, if G = \langle g \rangle where g has order n, then there exists an isomorphism \phi: \mathbb{Z}_n \to G defined by \phi(k) = g^k for k = 0, 1, \dots, n-1. This mapping preserves the group operation because \phi(k + m \mod n) = g^{k+m} = g^k g^m = \phi(k) \phi(m), and it is bijective since g generates G and has order n. Similarly, all infinite cyclic groups are isomorphic to the additive group of integers \mathbb{Z}. For an infinite cyclic group G = \langle a \rangle with a \neq e, the map \phi: \mathbb{Z} \to G given by \phi(k) = a^k is an isomorphism, as it is a bijective homomorphism: surjective because every element is a power of a, injective because the kernel is trivial (no nonzero k satisfies a^k = e), and it preserves addition via the group operation. The classification of cyclic groups up to is thus determined solely by their : there is exactly one isomorphism class for each finite n (namely, \mathbb{Z}_n) and one for the infinite case (\mathbb{Z}). No two cyclic groups of the same are non-, as the explicit isomorphisms above establish equivalence. This follows from the first isomorphism theorem applied to the natural surjection from \mathbb{Z} (or \mathbb{Z}_n) onto the , yielding the as the appropriate multiple of the . The first isomorphism theorem also applies directly to quotients of cyclic groups. For d \mid n, the subgroup \langle d \rangle of \mathbb{Z}_n has order n/d, and the quotient \mathbb{Z}_n / \langle d \rangle \cong \mathbb{Z}_{d}, as the natural projection map from \mathbb{Z}_n to \mathbb{Z}_{d} (reduction modulo d) has kernel \langle d \rangle. This isomorphism highlights how cyclic groups' structure simplifies under quotients by cyclic subgroups. To contrast, non-cyclic groups of the same order as a cyclic one are not isomorphic to it. For example, the \mathbb{Z}_2 \times \mathbb{Z}_2 has order 4 but is not cyclic, as all its non-identity elements have order 2, whereas \mathbb{Z}_4 has an element of order 4; thus, they are not isomorphic since isomorphisms preserve element orders. Specifically, \mathbb{Z}_2 \times \mathbb{Z}_2 has three elements of order 2, while \mathbb{Z}_4 has only one.

Automorphisms

An of a group G is a bijective \alpha: G \to G, preserving the group operation in both directions. The collection of all such automorphisms, denoted \Aut(G), forms a group under , where the is defined by (\alpha \circ \beta)(x) = \alpha(\beta(x)) for all x \in G, the is the map \id_G, and the of \alpha is \alpha^{-1}, which exists since \alpha is bijective. Automorphisms represent the symmetries of the group itself, capturing ways to relabel elements while maintaining the . A key subgroup of \Aut(G) consists of the inner automorphisms, generated by conjugation: for each g \in G, the map c_g: x \mapsto g x g^{-1} is an automorphism, and the set \Inn(G) = \{ c_g \mid g \in G \} forms a normal subgroup of \Aut(G). The conjugation action defines a homomorphism from G to \Aut(G) via g \mapsto c_g, with kernel equal to the center Z(G) = \{ z \in G \mid z x = x z \ \forall x \in G \}; by the first isomorphism theorem, this yields \Inn(G) \cong G / Z(G). Automorphisms outside \Inn(G), called outer automorphisms, exist when \Aut(G) / \Inn(G) is nontrivial, measuring "external" symmetries beyond conjugation. For the cyclic group \mathbb{Z}_n under modulo n, the automorphisms are by integers k coprime to n: \phi_k(a) = k a \mod n, and \Aut(\mathbb{Z}_n) \cong U(n), the of units modulo n. Here, the order of \Aut(\mathbb{Z}_n) is \phi(n), , illustrating how automorphisms quantify the invertible scalings preserving the cyclic structure. The holomorph of G, denoted \Hol(G) = G \rtimes \Aut(G), is the semidirect product where \Aut(G) acts on G by evaluation, providing a unified framework to study both the group elements and their symmetries as a single larger group.

References

  1. [1]
    3.3: Isomorphic Groups - Mathematics LibreTexts
    Sep 25, 2021 · Isomorphic groups have identical structures, though the elements of one group may differ greatly from those of the other.
  2. [2]
    [PDF] MATH 415 Modern Algebra I Lecture 15: Isomorphisms of groups.
    In other words, an isomorphism is a bijective homomorphism. The group G is said to be isomorphic to H if there exists an isomorphism f : G → H. Notation: G.
  3. [3]
    [PDF] Lecture 4.1: Homomorphisms and isomorphisms
    We will study a special type of function between groups, called a homomorphism. An isomorphism is a special type of homomorphism. The Greek roots “homo” and. “ ...
  4. [4]
    Group Isomorphism - an overview | ScienceDirect Topics
    From the point of view of abstract algebra, isomorphic groups have the same properties and they are indistinguishable. Two groups are isomorphic if there ...<|control11|><|separator|>
  5. [5]
    [PDF] Zeta Functions of Classical Groups and Class Two Nilpotent Groups
    Apr 2, 2019 · For instance, the primary problem in group theory is the classification of groups up to isomorphism. The enumeration of such classes of ...<|control11|><|separator|>
  6. [6]
    [PDF] homomorphisms.pdf - Keith Conrad
    A homomorphism is a function that preserves group operations, where f(xy) = f(x)f(y) for all x and y in G.Missing: phi| textbook
  7. [7]
    [PDF] ISOMORPHISMS 1. Introduction Groups that are not literally the ...
    If f : G → CG is an isomorphism its inverse function f−1 : CG → G is also an isomorphism. Proof. First we show f−1 is a homomorphism. For y ∈ CG, the definition ...
  8. [8]
    Isomorphism -- from Wolfram MathWorld
    Formally, an isomorphism is bijective morphism. Informally, an isomorphism is a map that preserves sets and relations among elements. " A is isomorphic to B " ...
  9. [9]
    [PDF] Group Theory - James Milne
    Mar 20, 2020 · ... (ISOMORPHISM THEOREM) Let H be a subgroup of G and N a normal ... Prove that if all Sylow subgroups of a finite group G are normal and abelian, ...
  10. [10]
    [PDF] GROUP THEORY NOTES FOR THE COURSE ALGEBRA 3, MATH ...
    We define its centralizer CG(H) to be the Dummit & Foote. §2.2 set {g ∈ G ... To give an action of a group G on a set A is equivalent to giving a homomorphism ...
  11. [11]
    [PDF] Chapter 1 Groups
    The group operation is matrix multiplication. GLn is called the general linear group. If G ≤ GLn(F) for some F and n then G is called a linear group of ...
  12. [12]
  13. [13]
    Abstract Algebra Second Edition - Northern Illinois University
    From this point on our book looks more like a traditional abstract algebra textbook. After rings we consider fields, and we include a discussion of root ...
  14. [14]
    None
    Summary of each segment:
  15. [15]
    [PDF] 12. Isomorphisms Look at the groups D3 and S3. They are clearly ...
    Isomorphisms. Look at the groups D3 and S3. They are clearly the same group. Given a symmetry of a triangle, the natural thing to do is to look at.
  16. [16]
    [PDF] Homework 4 solutions.
    Homework 4 solutions. Problem 7.4. Produce a specific isomorphism between S3 and D3. How many different isomorphisms are there from S3 to D3? Answer.
  17. [17]
    [PDF] On Groups That Are Isomorphic to a Proper Subgroup
    Clearly, no finite group can have this property, but what about Z, Q, R, or C, the familiar additive groups of integers, rational, real, and complex numbers?
  18. [18]
    [PDF] The 3 dimensional rotation group - Purdue Math
    matrices O(3) forms a subgroup of GL3(R). Let SO(3) = 1A 2 O(3) | det A = 1l. Lemma 13.2. SO(3) is a subgroup of O(3). Proof. If A, B 2 SO(3), then det(AB) ...
  19. [19]
    11.2: The Isomorphism Theorms - Mathematics LibreTexts
    Jun 4, 2022 · Theorem 11.10 . First Isomorphism Theorem. If ψ : G → H is a group homomorphism with K = ker ⁡ ψ , then K is normal in G . Let ϕ : G → G ...
  20. [20]
    [PDF] 18.703 Modern Algebra, The Isomorphism Theorems
    Theorem 10.1 (First Isomorphism Theorem). Let φ: G −→ G/ be a homomorphism of groups. Suppose that φ is onto and let H be the kernel of φ. Then G/ is ...
  21. [21]
    [PDF] The Isomorphism Theorems
    Sep 25, 2006 · The isomorphism theorems are based on a simple basic result on homo- morphisms. For a group G and N£G we let π : G −→ G/N be the projection.
  22. [22]
    [PDF] Cyclic groups - Purdue Math
    A subgroup of a cyclic group is cyclic. Proof. We may assume that the group is either Z or Zn. In the first case, we proved that any subgroup is Zd for some d.
  23. [23]
    [PDF] The First Isomorphism Theorem
    In this case the First Isomorphism Theorem then implies that Z/mZ ∼ = G. Thus: G = ⟨g⟩ ∼= ( Z if |g| = ∞, Z/mZ if |g| = m. This shows that, up to isomorphism, ...
  24. [24]
    None
    ### Summary of Proof: Klein 4-Group and Z4 Are Not Isomorphic
  25. [25]
    Group Automorphism -- from Wolfram MathWorld
    A group automorphism is an isomorphism from a group to itself. If G is a finite multiplicative group, an automorphism of G can be described as a way of ...
  26. [26]
    [PDF] Automorphisms
    Conjugation by g is called an inner automorphism of G and the subgroup of Aut(G) consisting of all inner automorphisms is denoted Inn(G). Kevin James.Missing: source | Show results with:source
  27. [27]
    G/Z(G) is Isomorphic to Inn(G) - Mathonline
    Proof: From proposition 1 we know that $G/Z(G)$ is isomorphic to $\mathrm{Inn}(G)$. So if $\mathrm{Inn}(G)$ is abelian then $G/Z(G)$ is abelian. But ...
  28. [28]
    [PDF] Lecture 4.6: Automorphisms - Mathematical and Statistical Sciences
    The automorphism group of Zn is Aut(Zn) = {σa | a ∈ U(n)} ∼= U(n), where σa : Zn −→ Zn , σa(1) = a . M. Macauley (Clemson). Lecture 4.6: Automorphisms. Math ...
  29. [29]
    holomorph of a group - PlanetMath.org
    Mar 22, 2013 · If G G is a group, the holomorph of G G is the group G⋊θAut(G) G ⋊ θ Aut ⁡ ( G ) under the identity map θ:Aut(G)→Aut(G) θ : Aut ⁡ ( G ) ...