Fact-checked by Grok 2 weeks ago

Isomorphism theorems

The isomorphism theorems are a collection of fundamental results in that establish isomorphisms between structures and homomorphic images in categories such as groups, rings, and modules, providing deep insights into the structural properties preserved under homomorphisms. These theorems, often numbering three or four depending on the context, generalize the idea that homomorphisms induce isomorphisms on objects by the , enabling the decomposition and classification of algebraic structures. In group theory, the first isomorphism theorem (also known as the fundamental homomorphism theorem) states that for any group homomorphism \phi: G \to G', the kernel \ker \phi is a normal subgroup of G, the image \operatorname{Im} \phi is a subgroup of G', and G / \ker \phi \cong \operatorname{Im} \phi, where the isomorphism is given by a (\ker \phi) \mapsto \phi(a). This theorem reduces the study of homomorphisms to isomorphisms between quotients and subgroups, highlighting how kernels capture the "loss of information" in mappings. The second isomorphism theorem, sometimes called the diamond theorem, applies to a subgroup H \leq G and a normal subgroup N \trianglelefteq G: it asserts that HN is a of G, H \cap N \trianglelefteq H, and HN / N \cong H / (H \cap N), with the isomorphism sending h (H \cap N) \mapsto hN. This result describes how products of subgroups interact with quotients, useful for analyzing subgroup lattices and extensions. The , or freshman theorem, addresses nested normal subgroups N \trianglelefteq H \trianglelefteq G: it shows that H / N \trianglelefteq G / N and (G / N) / (H / N) \cong G / H, via the map (aN) (H / N) \mapsto aH. This theorem justifies "cancelling" intermediate subgroups in quotient constructions, simplifying computations in group classifications. Finally, the fourth isomorphism theorem (or correspondence theorem) for a normal subgroup N \trianglelefteq G establishes a lattice isomorphism between the subgroups of G/N and the subgroups of G containing N, given by A/N \leftrightarrow A for N \leq A \leq G; moreover, A \trianglelefteq G if and only if A/N \trianglelefteq G/N. This bijection preserves normality and order, forming the backbone for induction on subgroup structures and applications in representation theory and cohomology. These theorems extend analogously to rings (where ideals play the role of normal subgroups) and modules (with submodules), underpinning much of and . Their proofs rely on the universal property of quotient maps and the first isomorphism theorem, making them cornerstones for understanding algebraic invariants like derived functors and exact sequences.

Introduction

Core Concepts

In , a is a structure-preserving between algebraic structures that maintains the operations defining those structures. For groups, a homomorphism \phi: G \to H between groups G and H is a function satisfying \phi(g_1 g_2) = \phi(g_1) \phi(g_2) for all g_1, g_2 \in G, where the operation in H is the group multiplication. For rings, a homomorphism \phi: R \to S between rings R and S preserves both addition and multiplication, so \phi(r_1 + r_2) = \phi(r_1) + \phi(r_2) and \phi(r_1 r_2) = \phi(r_1) \phi(r_2) for all r_1, r_2 \in R. An isomorphism is a bijective homomorphism, establishing an exact structural equivalence between the domains. The kernel of a homomorphism captures the elements that map to the trivial or zero element in the codomain, forming a key invariant. For a group homomorphism \phi: G \to H, the kernel \ker \phi = \{ g \in G \mid \phi(g) = e_H \}, where e_H is the identity in H, is always a normal subgroup of G. In the ring setting, for \phi: R \to S, the kernel \ker \phi = \{ r \in R \mid \phi(r) = 0_S \}, where $0_S is the zero in S, is an ideal of R. This kernel measures the "loss of information" in the mapping and enables the construction of quotient structures. The image of a homomorphism consists of the elements in the codomain that are actually reached by the map, generating a substructure. For \phi: G \to H, the image \operatorname{im} \phi = \{ \phi(g) \mid g \in G \} is a subgroup of H. Similarly, for a ring homomorphism \phi: R \to S, \operatorname{im} \phi is a subring of S. The image highlights the range of the structure-preserving action. Quotient structures arise from kernels by partitioning the domain into equivalence classes, or cosets in groups and residue classes in rings, to form a new algebraic object. In groups, if N = \ker \phi is a of G, the G/N consists of the left cosets \{ gN \mid g \in G \} with multiplication (g_1 N)(g_2 N) = (g_1 g_2) N. For rings, if I = \ker \phi is an of R, the R/I has elements as cosets r + I with operations (r_1 + I) + (r_2 + I) = (r_1 + r_2) + I and (r_1 + I)(r_2 + I) = (r_1 r_2) + I. These quotients effectively "mod out" the kernel to simplify the structure while preserving essential properties. A concrete illustration is the canonical homomorphism \phi: \mathbb{Z} \to \mathbb{Z}/n\mathbb{Z} defined by \phi(k) = k + n\mathbb{Z}, where \mathbb{Z}/n\mathbb{Z} is the of integers n. Here, the is n\mathbb{Z} = \{ kn \mid k \in \mathbb{Z} \}, the multiples of n, which partitions \mathbb{Z} into n cosets forming the \mathbb{Z}/n\mathbb{Z}.

Significance in Algebra

The isomorphism theorems establish natural isomorphisms between quotient structures derived from and the images of those homomorphisms, ensuring that algebraic properties such as binary operations and identities are preserved across these equivalent forms. This equivalence simplifies the analysis of algebraic objects by allowing computations in more manageable quotient settings without altering the intrinsic structure. For instance, in group theory, the of a homomorphism serves as a , enabling the to mirror the image faithfully. In the of algebraic structures, these theorems are because they determine when two groups, rings, or modules are , treating them as equivalent up to relabeling of elements and thus reducing the problem to identifying distinct isomorphism classes. This approach underpins major efforts, such as the of finite abelian groups or the structure theorem for finitely generated modules over domains, where uniqueness holds . By focusing on invariants preserved under these theorems, mathematicians can catalog structures systematically. The theorems enhance proof techniques in by permitting the reduction of complex problems to quotients, where solutions—such as finding elements satisfying certain relations—can be lifted back to the original structure via the . Additionally, they connect to theory, where subgroups or ideals ordered by inclusion form a , and the correspondence theorem induces an between the containing a fixed and the in the quotient, preserving meets and joins. This structure reveals hierarchical relationships among subobjects./07%3A_Homomorphisms_and_the_Isomorphism_Theorems/7.02%3A_The_Isomorphism_Theorems) A key example of their significance is in the Jordan-Hölder theorem, which asserts the uniqueness of composition factors in a composition series of a group; this result depends on the isomorphism theorems to refine series while maintaining isomorphic factor groups, ensuring the multiset of simple factors is invariant regardless of the series chosen. This invariance facilitates the decomposition of groups into irreducible components, advancing deeper structural insights.

Historical Development

Origins and Early Formulations

The origins of the isomorphism theorems trace back to the mid-19th century developments in group theory, particularly Arthur Cayley's pioneering work on substitution groups. In the 1850s, Cayley explored the structure of groups through permutations and implicitly employed ideas akin to quotients when analyzing the relations among group elements, such as in his studies of finite groups generated by symbols satisfying certain equations. His 1854 paper provided the first abstract definition of a group, emphasizing combinatorial laws that laid groundwork for recognizing structural equivalences later formalized as isomorphisms. Felix Klein's , introduced in , further advanced these concepts by classifying geometries through their associated transformation groups, highlighting the preservation of structure under group actions as a core principle. This approach influenced the understanding of isomorphisms as mappings that maintain invariant properties, bridging concrete permutation examples with broader abstract notions of in group structures. By the late , the transition from concrete permutation groups to formulations gained momentum, with explicit statements of principles appearing in key texts around 1900. Heinrich Weber's 1895 Lehrbuch der Algebra discussed the equivalence of isomorphic groups as essentially identical mathematical objects, incorporating and ideas in the context of finite groups. Similarly, William Burnside's 1897 Theory of Groups of Finite Order examined isomorphisms between groups and introduced related theorems on subgroup structures, solidifying these concepts in systematic group classification. Emmy Noether's 1921 paper Idealtheorie in Ringbereichen served as a foundational work that crystallized isomorphism theorems in a general algebraic framework, building directly on these earlier group-theoretic roots. While her contributions extended to rings, they underscored the abstract power of these theorems originating from group theory's evolution.

Evolution Across Structures

In the early 20th century, the isomorphism theorems, initially formulated for groups, were extended to ring theory by Emmy Noether in her seminal 1921 paper "Idealtheorie in Ringbereichen." Noether unified the concepts of ideals and quotient rings, drawing direct analogies to normal subgroups and quotient groups from group theory, which allowed for the establishment of corresponding isomorphism theorems that describe the structure of homomorphic images and kernels in rings. This extension was motivated by the need to generalize algebraic structures beyond groups to handle commutative and non-commutative rings systematically, providing a foundation for modern commutative algebra. Building on her ring theory work, Noether advanced the framework to modules around , introducing modules over rings as a generalization of both ideals and spaces, and adapting the isomorphism theorems to capture relationships between submodules, quotients, and homomorphisms. This was driven by applications in and non-commutative algebras, where modules offered a versatile tool for studying linear actions. In the 1950s, and further refined these ideas in their foundational text on , integrating module isomorphism theorems into a broader categorical to support derived functors and exact sequences in homological contexts. The generalization to universal algebra was pioneered by in 1935, who extended the theorems to arbitrary algebraic structures defined by operations and equations, encompassing varieties of algebras closed under homomorphic images, subalgebras, and products. Birkhoff's approach was motivated by the desire to classify abstract algebras uniformly, beyond specific cases like groups or rings, facilitating the study of equational classes. From the 1950s onward, , as developed by and , profoundly influenced this evolution by abstracting the theorems to apply to functors between categories and natural transformations, emphasizing universal properties and enabling applications across diverse mathematical domains. A pivotal role in standardizing these extensions across structures was played by Bartel Leendert van der Waerden's "Moderne Algebra," first published in 1930, which introduced consistent notations and axiomatic treatments for groups, rings, and fields, thereby disseminating the unified framework to a wider audience and influencing subsequent algebraic developments.

Theorems in Group Theory

First Isomorphism Theorem

The first isomorphism theorem for groups, also known as the fundamental homomorphism theorem, states that if \phi: G \to H is a , then \ker \phi is a of G, \operatorname{im} \phi is a of H, and G / \ker \phi \cong \operatorname{im} \phi as groups. The isomorphism is induced by \phi, sending the g \ker \phi to \phi(g). To prove the theorem, define a map \psi: G / \ker \phi \to \operatorname{im} \phi by \psi(g \ker \phi) = \phi(g) for all g \in G. This map is well-defined: if g' \ker \phi = g \ker \phi, then g' g^{-1} \in \ker \phi, so \phi(g') = \phi(g). Moreover, \psi preserves the group operation, since \psi((g_1 \ker \phi)(g_2 \ker \phi)) = \psi(g_1 g_2 \ker \phi) = \phi(g_1 g_2) = \phi(g_1) \phi(g_2) = \psi(g_1 \ker \phi) \psi(g_2 \ker \phi). Thus, \psi is a . It is injective because if \psi(g \ker \phi) = e_H, then \phi(g) = e_H, so g \in \ker \phi and g \ker \phi = \ker \phi, the identity in the quotient. It is surjective since every element of \operatorname{im} \phi is \phi(g) for some g \in G, which equals \psi(g \ker \phi). Therefore, \psi is an . A concrete example is the \phi: \mathbb{Z} \to \mathbb{Z}/n\mathbb{Z} defined by \phi(k) = k \mod n. Here, \ker \phi = n\mathbb{Z}, a , and \operatorname{im} \phi = \mathbb{Z}/n\mathbb{Z}. The \mathbb{Z}/n\mathbb{Z} \cong \mathbb{Z}/n\mathbb{Z}, confirming the . This illustrates how the theorem classifies cyclic groups via homomorphisms. This theorem underpins the $1 \to \ker \phi \to G \xrightarrow{\phi} \operatorname{im} \phi \to 1, exact by construction, where the kernel measures the failure of injectivity.

Second Isomorphism Theorem

The second isomorphism theorem for groups states that if H is a of G and N \trianglelefteq G is a , then HN is a of G, H \cap N \trianglelefteq H, and HN / N \cong H / (H \cap N) as groups. The sends h (H \cap N) to hN. To prove this, consider the restriction of the natural projection \pi: G \to G/N to H, yielding a f: H \to G/N by f(h) = hN. The of f is H \cap N, since f(h) = N (the ) if and only if h \in N. The image of f is HN/N, as elements are hN for h \in H. By the first theorem, H / (H \cap N) \cong HN / N. The explicit isomorphism \phi: H / (H \cap N) \to HN / N is \phi(h (H \cap N)) = hN, well-defined because if h' (H \cap N) = h (H \cap N), then h' h^{-1} \in H \cap N \subseteq N, so h'N = hN. This map is a , injective (kernel trivial), and surjective (every h'N = (h' n')N = h'N for n' \in N). A concrete example is G = S_3, H = \langle (1\ 2) \rangle, N = A_3 = \langle (1\ 2\ 3) \rangle. Then H \cap N = \{e\}, HN = S_3, so S_3 / A_3 \cong H / \{e\} \cong \mathbb{Z}/2\mathbb{Z}, matching the order-2 quotient. If H \cap N = \{e\}, then HN = H \times N if H normalizes N, but generally a .

Third Isomorphism Theorem

The third isomorphism theorem for groups states that if N \trianglelefteq H \trianglelefteq G with normal subgroups, then H/N \trianglelefteq G/N and (G/N) / (H/N) \cong G/H as groups. The isomorphism sends (gN) (H/N) to gH. To prove this, consider the natural surjective homomorphism \phi: G/N \to G/H by \phi(gN) = gH. This is well-defined: if gN = g'N, then g'^{-1} g \in N \subseteq H, so gH = g'H. It preserves the group operation since \phi((g_1 N)(g_2 N)) = \phi(g_1 g_2 N) = g_1 g_2 H = (g_1 H)(g_2 H) = \phi(g_1 N) \phi(g_2 N), and is surjective as every gH is hit by gN. The kernel is \{gN \mid g \in H\} = H/N, a normal subgroup of G/N. By the first isomorphism theorem, (G/N) / (H/N) \cong G/H. The explicit isomorphism sends the coset (gN) (H/N) to gH, bijective and preserving the group structure. For an example, take G = \mathbb{Z}, H = 2\mathbb{Z}, N = 4\mathbb{Z}, all normal since abelian. Then G/N = \mathbb{Z}/4\mathbb{Z}, H/N = 2\mathbb{Z}/4\mathbb{Z} \cong \mathbb{Z}/2\mathbb{Z}, so (G/N)/(H/N) \cong \mathbb{Z}/2\mathbb{Z}, and G/H = \mathbb{Z}/2\mathbb{Z}, verifying the isomorphism. This simplifies computations in abelian group classifications. This theorem relates to , where successive quotients are preserved under intermediate normals.

Correspondence Theorem

The , also known as the or sometimes the fourth , establishes a bijective between the of a group G that contain a fixed N \trianglelefteq G and the of the G/N. Specifically, the \phi: H \mapsto H/N for N \leq H \leq G is a from the set of such H to the set of all of G/N, and the inverse is given by K/N \mapsto \pi^{-1}(K), where \pi: G \to G/N is the canonical projection and K \leq G/N. This preserves the structure: if H_1, H_2 are containing N, then \phi(H_1 \cap H_2) = \phi(H_1) \cap \phi(H_2) and \phi(H_1 H_2) = \phi(H_1) \phi(H_2). Moreover, H \trianglelefteq G \phi(H) \trianglelefteq G/N, ensuring that correspond to . To prove the theorem, first verify that \phi is well-defined: for N \leq H \leq G, H/N is indeed a of G/N by the properties of groups. The map is injective because if \phi(H_1) = \phi(H_2), then H_1/N = H_2/N, implying H_1 = H_2 since the natural map restricts to an on cosets. Surjectivity follows from the fact that for any K \leq G/N, the preimage \pi^{-1}(K) is a of G containing N, and \phi(\pi^{-1}(K)) = K. Preservation of intersections and products arises from the corresponding properties in groups: for intersections, \pi(H_1 \cap H_2) = \pi(H_1) \cap \pi(H_2), and similarly for products using the second theorem indirectly. Normality preservation holds because conjugates in the lift appropriately via the . This bijection thus induces a between the poset of s containing N ordered by inclusion and the of G/N. The structure emphasized by the theorem reveals how the of G above N mirrors that of G/N, providing a tool to study relations in quotients without recomputing from scratch. This is particularly useful in classifying groups or analyzing , as successive quotients inherit properties. The naming as the "fourth isomorphism theorem" arises in some texts to distinguish it from the first three, which focus on images and kernels of homomorphisms, though the correspondence theorem generalizes these by considering all intermediate s rather than just successive ones; variations in numbering reflect different pedagogical emphases. A concrete example illustrates the full lattice correspondence in the V_4 = \mathbb{Z}/2\mathbb{Z} \times \mathbb{Z}/2\mathbb{Z} = \langle a, b \mid a^2 = b^2 = (ab)^2 = e \rangle, taking the trivial N = \{e\}. Here, V_4/N \cong V_4, and the correspondence \phi: H \mapsto H is the identity map on the lattice. The proper nontrivial subgroups of V_4 are \langle a \rangle = \{e, a\}, \langle b \rangle = \{e, b\}, and \langle ab \rangle = \{e, ab\}, all since V_4 is elian. The lattice has these three subgroups at the "middle" level, with intersections pairwise trivial and products yielding the whole group, directly mirrored in the quotient lattice. This trivial case nonetheless demonstrates the theorem's preservation of the complete lattice structure inherent to elementary elian 2-groups.

Theorems in Ring Theory

First Isomorphism Theorem

The first isomorphism theorem for rings states that if \phi: R \to S is a between rings R and S, then \ker \phi is an of R, \operatorname{im} \phi is a of S, and the R / \ker \phi is isomorphic to \operatorname{im} \phi as rings. This theorem mirrors the group version but accounts for the multiplicative of rings. To prove the theorem, first note that \ker \phi = \{ r \in R \mid \phi(r) = 0 \} is an ideal: it absorbs multiplication from R on both sides, as \phi(r a) = \phi(r) \phi(a) = 0 and \phi(a r) = \phi(a) \phi(r) = 0 for r \in \ker \phi, a \in R. The map \psi: R / \ker \phi \to \operatorname{im} \phi defined by \psi(r + \ker \phi) = \phi(r) is well-defined, since if r - r' \in \ker \phi, then \phi(r) = \phi(r'). It preserves addition: \psi((r_1 + \ker \phi) + (r_2 + \ker \phi)) = \phi(r_1 + r_2) = \phi(r_1) + \phi(r_2). It preserves multiplication: \psi((r_1 + \ker \phi)(r_2 + \ker \phi)) = \psi(r_1 r_2 + \ker \phi) = \phi(r_1 r_2) = \phi(r_1) \phi(r_2). It maps the identity coset to the identity in \operatorname{im} \phi. Thus, \psi is a ring homomorphism. Injectivity follows if \psi(r + \ker \phi) = 0, then \phi(r) = 0, so r \in \ker \phi. Surjectivity holds as every element in \operatorname{im} \phi is \phi(r) for some r \in R. Hence, \psi is a ring isomorphism. A concrete example is the projection homomorphism \phi: \mathbb{Z} \to \mathbb{Z}/n\mathbb{Z} defined by \phi(k) = k + n\mathbb{Z}. Here, \ker \phi = n\mathbb{Z}, an ideal, and \operatorname{im} \phi = \mathbb{Z}/n\mathbb{Z}, so \mathbb{Z}/n\mathbb{Z} \cong \mathbb{Z}/n\mathbb{Z}, illustrating the theorem tautologically but confirming the structure. More non-trivially, the evaluation map \phi: \mathbb{Z} \to \mathbb{Z} sending p(x) to p(0) has kernel the principal ideal (x), yielding \mathbb{Z}/(x) \cong \mathbb{Z}. This theorem supports exact sequences in , such as $0 \to \ker \phi \to R \xrightarrow{\phi} \operatorname{im} \phi \to 0, emphasizing as kernels of maps.

Second Isomorphism Theorem

The second isomorphism theorem for rings states that if S is a of a R and I is an of R, then S + I = \{ s + a \mid s \in S, a \in I \} is a of R, S \cap I is an of S, and (S + I)/I \cong S / (S \cap I) as rings. To prove this, S + I is a subring: it is closed under addition and contains additive inverses, and for multiplication, (s_1 + a_1)(s_2 + a_2) = s_1 s_2 + s_1 a_2 + a_1 s_2 + a_1 a_2 \in S + I since s_1 s_2 \in S, a_1 a_2 \in I, and s_1 a_2, a_1 s_2 \in I by ideal absorption. Similarly, S \cap I is an ideal in S. Consider the projection \pi: R \to R/I, restricted to S giving f: S \to R/I by f(s) = s + I. The kernel is S \cap I, and the image is (S + I)/I. By the first isomorphism theorem, S / (S \cap I) \cong (S + I)/I. The explicit isomorphism \phi: S / (S \cap I) \to (S + I)/I by \phi(s + (S \cap I)) = s + I is a ring homomorphism: well-defined (differences in kernel map to same coset), preserves operations, and is bijective. A concrete example is R = \mathbb{Z}, S = \mathbb{Z}, I = (x). Then S \cap I = 0, S + I = \mathbb{Z}, so \mathbb{Z}/(x) \cong \mathbb{Z}/0 \cong \mathbb{Z}, matching the earlier evaluation map. Another: In \mathbb{Z}, take S = 2\mathbb{Z} (as additive subgroup, but noting unity issues in some conventions), I = 3\mathbb{Z}; then S \cap I = 6\mathbb{Z}, S + I = \mathbb{Z}, yielding \mathbb{Z}/3\mathbb{Z} \cong 2\mathbb{Z}/6\mathbb{Z} \cong \mathbb{Z}/3\mathbb{Z}, both cyclic of order 3. If S \cap I = 0, then S \cong (S + I)/I, and under suitable conditions, S + I = S \oplus I additively.

Third Isomorphism Theorem

The third isomorphism theorem for rings states that if J \subseteq I are ideals of a ring R, then I/J is an ideal of R/J, and (R/J) / (I/J) \cong R/I as rings. This allows simplifying nested quotient constructions. To prove this, the map \phi: R/J \to R/I defined by \phi(r + J) = r + I is well-defined: if r + J = r' + J, then r - r' \in J \subseteq I, so r + I = r' + I. It is a ring homomorphism: \phi((r_1 + J) + (r_2 + J)) = (r_1 + r_2) + I = \phi(r_1 + J) + \phi(r_2 + J), and similarly for multiplication \phi((r_1 + J)(r_2 + J)) = r_1 r_2 + I = (r_1 + I)(r_2 + I). It is surjective as every r + I is hit. The kernel is \{ r + J \mid r \in I \} = I/J. By the first isomorphism theorem, (R/J) / (I/J) \cong R/I. The explicit isomorphism sends (r + J) + (I/J) to r + I. For an example, take R = \mathbb{Z}, J = 4\mathbb{Z}, I = 2\mathbb{Z}. Then R/J = \mathbb{Z}/4\mathbb{Z}, I/J = 2\mathbb{Z}/4\mathbb{Z} \cong \mathbb{Z}/2\mathbb{Z} (generated by $2 + 4\mathbb{Z}), and (\mathbb{Z}/4\mathbb{Z}) / (2\mathbb{Z}/4\mathbb{Z}) \cong \mathbb{Z}/2\mathbb{Z}, matching R/I = \mathbb{Z}/2\mathbb{Z}. This theorem aids in homological algebra for rings, linking to exact sequences like $0 \to I/J \to R/J \to R/I \to 0.

Ideal Lattice Structure

In ring theory, the correspondence theorem provides a lattice isomorphism between the ideals of a ring R that contain a fixed ideal I and the ideals of the quotient ring R/I. Specifically, the map sending each ideal K with I \leq K \leq R to the ideal K/I in R/I is a bijection that preserves the lattice operations: for ideals K_1, K_2 containing I, (K_1 \cap K_2)/I = K_1/I \cap K_2/I and (K_1 + K_2)/I = K_1/I + K_2/I, where + denotes the ideal sum. In the commutative case, the sum coincides with the ideal product, ensuring the structure is preserved under multiplication as well. The proof establishes this bijection by considering the natural projection \pi: R \to R/I. The map K \mapsto K/I is well-defined since I \subseteq K implies K/I is an ideal in R/I; it is injective because if K_1/I = K_2/I, then K_1 = K_2; and it is surjective because any ideal J of R/I arises as J = \pi^{-1}(J)/I, where \pi^{-1}(J) is an ideal of R containing I. The preservation of lattice operations follows directly: intersections pull back under \pi and sums project forward compatibly. Equivalently, this bijection arises from ring homomorphisms out of R/I, as every ideal of R/I is the kernel of some ring homomorphism \phi: R/I \to S for a ring S, corresponding to the preimage ideal \pi^{-1}(\ker \phi) in R containing I. A concrete example illustrates this structure in polynomial rings. Consider R = \mathbb{C}[x,y] and I = (x). The quotient R/I \cong \mathbb{C} is a principal ideal domain, so its ideals are of the form (f(y)) for polynomials f \in \mathbb{C}. Under the correspondence, these map to ideals (x, f(y)) in R, which properly contain I and form a chain ordered by divisibility of the generators. For instance, the zero ideal in R/I corresponds to I itself, while the unit ideal corresponds to R. This correspondence extends to special classes of ideals in quotients. Prime ideals of R/I biject with prime ideals of R containing I, since R/I is an if and only if I is prime; similarly, maximal ideals of R/I correspond to maximal ideals containing I, ensuring R/I is a precisely when I is maximal. These variations underpin key results in , such as the lying-over theorem in integral extensions. Geometrically, this structure relates to the of rings: the prime ideals of R/I form the closed V(I) = \{\mathfrak{p} \in \operatorname{Spec} R \mid I \subseteq \mathfrak{p}\} of \operatorname{Spec} R, intuitively representing the subvariety defined by the equations in I within the affine space associated to R. This provides a bridge to , where quotients model restrictions to subschemes.

Theorems in Module Theory

First Isomorphism Theorem

The first isomorphism theorem for modules asserts that for any R and R-modules M and N, if \phi: M \to N is a of R-modules, then the module M / \ker(\phi) is isomorphic to the \operatorname{im}(\phi) as R-modules. This result generalizes the corresponding theorem in group and to the more flexible setting of modules, where by ring elements plays a central role. To prove the theorem, define a map \psi: M / \ker(\phi) \to \operatorname{im}(\phi) by \psi(m + \ker(\phi)) = \phi(m) for all m \in M. This map is well-defined: if m' + \ker(\phi) = m + \ker(\phi), then m' - m \in \ker(\phi), so \phi(m') = \phi(m). Moreover, \psi preserves addition, since \psi((m_1 + \ker(\phi)) + (m_2 + \ker(\phi))) = \phi(m_1 + m_2) = \phi(m_1) + \phi(m_2) = \psi(m_1 + \ker(\phi)) + \psi(m_2 + \ker(\phi)), and it respects , as \psi(r(m + \ker(\phi))) = \phi(rm) = r\phi(m) = r \psi(m + \ker(\phi)) for r \in [R](/page/R). Thus, \psi is a of R-modules. It is injective because if \psi(m + \ker(\phi)) = 0, then \phi(m) = 0, so m \in \ker(\phi) and m + \ker(\phi) = 0 in the . It is surjective since every element of \operatorname{im}(\phi) is \phi(m) for some m \in M, which equals \psi(m + \ker(\phi)). Therefore, \psi is an . A concrete example illustrates the theorem: consider the \mathbb{R}-module homomorphism \phi: \mathbb{R}^2 \to \mathbb{R} defined by \phi(x, y) = x. Here, \ker(\phi) = \{(0, y) \mid y \in \mathbb{R}\}, which is a one-dimensional subspace isomorphic to \mathbb{R}. The quotient \mathbb{R}^2 / \ker(\phi) consists of cosets (x, y) + \ker(\phi), which can be identified with the x-coordinate, yielding an isomorphism to \mathbb{R}; indeed, \operatorname{im}(\phi) = \mathbb{R}. This theorem underpins the short exact sequence $0 \to \ker(\phi) \to M \xrightarrow{\phi} \operatorname{im}(\phi) \to 0, which is exact by construction, highlighting the kernel as the obstruction to injectivity and the image as the effective range of the map.

Second Isomorphism Theorem

The second isomorphism theorem for modules states that if M is an R-module and A, B are submodules of M, then A + B and A \cap B are also submodules of M, and there is an isomorphism of R-modules \frac{A + B}{B} \cong \frac{A}{A \cap B}. To prove this, consider the natural projection map \pi: M \to M/B, which is an R-. Restrict \pi to the submodule A to obtain a f: A \to M/B defined by f(a) = a + B. The of f is A \cap B, since f(a) = 0 a \in B. The image of f is (A + B)/B, as every element in the image is of the form a + B for a \in A, and any such can be written as (a + b') + B for b' \in B. By the first for modules, A / (A \cap B) \cong (A + B)/B. The explicit isomorphism is given by the map \phi: A / (A \cap B) \to (A + B)/B defined by \phi(a + (A \cap B)) = a + B, which is well-defined because if a' - a \in A \cap B \subseteq B, then a' + B = a + B. This map is an R-module homomorphism, injective (since the kernel is trivial), and surjective (as every coset in the image arises from elements in A). A concrete example arises in the \mathbb{Z}-module \mathbb{Z}, taking A = 2\mathbb{Z} and B = 3\mathbb{Z}. Here, A \cap B = 6\mathbb{Z} and A + B = \mathbb{Z}, so the theorem yields \frac{\mathbb{Z}}{3\mathbb{Z}} \cong \frac{2\mathbb{Z}}{6\mathbb{Z}}. Both sides are cyclic groups of order 3, confirming the isomorphism. If A \cap B = 0, then the theorem implies A \cong (A + B)/B, and in this case A + B = A \oplus B is a direct sum of submodules.

Third Isomorphism Theorem

The third isomorphism theorem for modules asserts that if N \leq M \leq P are submodules of an R-module P, where R is a ring, then there is a canonical isomorphism of R-modules (P/N) / (M/N) \cong P/M. This result refines the structure of successive quotients by submodules, showing that the quotient of the larger quotient by the intermediate one yields the direct quotient by the intermediate submodule. To prove this, consider the natural surjective module homomorphism \phi: P/N \to P/M defined by \phi(p + N) = p + M for all p \in P. This map is well-defined: if p + N = p' + N, then p - p' \in N \subseteq M, so p + M = p' + M. It preserves addition and scalar multiplication since both operations are induced from P, and it is surjective because every coset in P/M is hit by elements from P. The kernel of \phi is precisely \{ p + N \mid p \in M \} = M/N, a submodule of P/N. By the first isomorphism theorem for modules, (P/N) / \ker \phi \cong P/M, which gives the desired isomorphism. The explicit isomorphism sends the coset (p + N) + (M/N) in (P/N)/(M/N) to p + M in P/M. This correspondence is bijective and respects the module structure, confirming the structural equivalence of these quotients. For a concrete illustration in the case of vector spaces (free modules over a ), take P = \mathbb{R}^3 with the standard basis \{e_1, e_2, e_3\}, N = \operatorname{[span](/page/Span)}\{e_1\} (the x-axis), and M = \operatorname{[span](/page/Span)}\{e_1, e_2\} (the xy-plane). Then P/N \cong \mathbb{R}^2 (the yz-plane), and M/N \cong \mathbb{R} (the y-direction in that plane), so (P/N)/(M/N) \cong \mathbb{R}. Similarly, P/M \cong \mathbb{R} (the z-axis), verifying the . This theorem connects to homological algebra through exact sequences: the short exact sequence $0 \to M/N \to P/N \to (P/N)/(M/N) \to 0 is exact, and the isomorphism identifies the final quotient with P/M, facilitating computations in derived functors and module resolutions.

Submodule Lattice Properties

In the context of module theory, the correspondence theorem establishes a bijection between the submodules L of an R-module M satisfying N \leq L \leq M, where N is a fixed submodule, and the submodules of the quotient module M/N. This bijection is given by L \mapsto L/N, and it preserves the lattice structure by mapping joins (submodule sums) to joins and meets (intersections) to meets: specifically, (L_1 + L_2)/N = L_1/N + L_2/N and (L_1 \cap L_2)/N = L_1/N \cap L_2/N. The proof of this correspondence follows from the universal property of quotient modules. The quotient map \pi: M \to M/N is a surjective R-module homomorphism with kernel N. For any submodule L with N \leq L \leq M, L/N is a submodule of M/N because it is the image of L under \pi and closed under the module operations. The map is injective since if L_1/N = L_2/N, then L_1 = L_2 by the second isomorphism theorem applied to the inclusions. Surjectivity holds because for any submodule K \subseteq M/N, the preimage \pi^{-1}(K) is a submodule of M containing N, and \pi(\pi^{-1}(K))/N = K. Preservation of lattice operations arises directly from the definitions of sums and intersections in quotient modules, ensuring the bijection is a lattice isomorphism. For example, consider the free \mathbb{Z}- \mathbb{Z}^n. The of its submodules forms a modular , satisfying the modular law (L_1 + L_2) \cap L_3 = L_1 + (L_2 \cap L_3) whenever L_2 \subseteq L_3, which aligns with the general modularity of submodule under the correspondence theorem. In special cases, such as Noetherian modules, the submodule satisfies the ascending chain condition (), meaning every ascending chain of submodules stabilizes after finitely many steps. The correspondence theorem implies that if M is Noetherian, then both the lattice of submodules containing N and the lattice of submodules of M/N satisfy . Similarly, for Artinian modules, the descending chain condition () holds, and the bijection preserves this property, ensuring no infinite strictly descending chains in the corresponding . Unlike the in , which benefits from the ring's multiplicative enabling concepts like prime or maximal , the submodule of a lacks inherent multiplication between elements, focusing solely on the additive and aspects without additional algebraic operations influencing the properties.

Generalizations

Universal Algebra Framework

In , the isomorphism theorems generalize to arbitrary algebras equipped with finitary operations, where homomorphisms preserve these operations, and structural quotients are formed by congruences—equivalence relations compatible with the operations. These theorems hold within varieties, which are equationally defined classes of algebras closed under homomorphic images (H), subalgebras (S), and arbitrary products (P), as established by Birkhoff's HSP theorem. Congruences on an algebra A, denoted \Con A, form an algebraic , enabling the lattice-theoretic underlying the theorems. The first isomorphism theorem states that for a homomorphism f: A \to B between algebras in a variety, the quotient algebra A / \ker(f) is isomorphic to the image \im(f), where \ker(f) is the kernel congruence \{(a, a') \in A \times A \mid f(a) = f(a')\}. This isomorphism arises from the natural surjection \nu: A \to A / \ker(f) composed with a bijection induced by f, confirming that the kernel fully captures the homomorphism's structural information. The second isomorphism theorem addresses interactions between subalgebras and congruences: if S is a of A and C \in \Con A, then S / (C \upharpoonright S) \cong SC / C, where C \upharpoonright S = C \cap (S \times S) is the restriction of C to S, and SC is the subalgebra generated by S and the elements of C. The third theorem extends this to nested congruences: if \theta \subseteq \phi in \Con A, then (A / \theta) / (\phi / \theta) \cong A / \phi, with \phi / \theta = \{([\bar{a}]_\theta, [\bar{b}]_\theta) \mid (\bar{a}, \bar{b}) \in \phi\}. Accompanying these is the congruence correspondence theorem, which establishes a between the congruences on A containing \theta and the congruences on the quotient A / \theta, preserving joins and meets. These theorems unify the specific cases in group, ring, and module theory as instances within varieties like groups or rings. For example, in the variety of Boolean algebras, congruences correspond to ideals, and the theorems describe how quotient Boolean algebras inherit properties from subalgebras and principal ideals generated by atoms. Similarly, in lattice varieties, the isomorphism theorems facilitate the study of sublattice quotients by deductive systems, which act as congruences.

Extensions to Other Categories

The isomorphism theorems extend naturally to the setting of abelian categories, where the structural properties of kernels, cokernels, and exact sequences allow for direct generalizations without relying on concrete elements. In an abelian category \mathcal{A}, the first isomorphism theorem asserts that for any morphism f: A \to B, the induced map A / \ker f \to \operatorname{im} f is an isomorphism, while the second and third theorems follow from the exactness of short exact sequences and the uniqueness of factorizations. These hold abstractly due to the abelian category axioms, which ensure that every monomorphism is the kernel of its cokernel and vice versa. A key tool in this context is the , which constructs long exact sequences from with exact rows and columns, thereby implying the isomorphism theorems for and within those sequences. Specifically, given a \begin{CD} 0 @>>> A @>f>> B @>g>> C @>>> 0 \\ @. @VV\alpha V @VV\beta V @VV\gamma V @. \\ 0 @>>> A' @>>f'> B' @>>g'> C' @>>> 0 \end{CD} with exact rows, the snake lemma yields an exact sequence \ker \alpha \to \ker \beta \to \ker \gamma \to \coker \alpha \to \coker \beta \to \coker \gamma, where the homomorphism links the cokernel of the first vertical map to the kernel of the last. This lemma underpins much of in abelian categories, including the long exact sequences in . Higher analogs of the isomorphism theorems appear in the five-lemma and nine-lemma, which provide criteria for morphisms to be isomorphisms in diagrams of exact sequences. The five-lemma states that in a commutative diagram of abelian groups (or objects in an abelian category) with exact rows, \begin{CD} A_1 @>>> A_2 @>>> A_3 @>>> A_4 @>>> A_5 \\ @VVf_1 V @VVf_2 V @VVf_3 V @VVf_4 V @VVf_5 V \\ B_1 @>>> B_2 @>>> B_3 @>>> B_4 @>>> B_5 \end{CD} if f_1 and f_5 are isomorphisms and the rows are exact, then f_3 is an isomorphism; variants cover cases where f_2 and f_4 are isomorphisms implying f_3 is. The nine-lemma extends this to a 3-by-3 grid of exact sequences, concluding that if the boundary morphisms are isomorphisms, the center morphism is as well, generalizing the third isomorphism theorem to longer complexes. These lemmas are essential for proving isomorphisms in derived functors and Ext groups. In triangulated categories, such as derived categories of abelian categories, the third isomorphism theorem finds a generalization through Verdier quotients. For a triangulated category \mathcal{T} and a thick triangulated subcategory \mathcal{A}, the Verdier quotient \mathcal{T}/\mathcal{A} is the localization of \mathcal{T} at the multiplicative system of morphisms whose cones lie in \mathcal{A}, yielding a triangulated category where objects of \mathcal{A} become zero and the quotient functor induces isomorphisms analogous to G/N \cong (G/M)/(N/M) for normal subgroups. This construction preserves distinguished triangles and shift functors, enabling the study of quotients in homological contexts like sheaf cohomology. A concrete example arises in the category of abelian groups \operatorname{Ab}, which is abelian and serves as the foundational setting for homological algebra; here, the isomorphism theorems apply directly to compute homology groups via chain complexes, with the snake lemma producing the connecting homomorphisms in the long exact sequence of a short exact sequence of complexes.

References

  1. [1]
    [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.
  2. [2]
    [PDF] Math 3230 Abstract Algebra I Sec 4.5: Isomorphism theorems
    The Fundamental Homomorphism Theorem (FHT) is the first of four basic theorems about homomorphism and their structure. These are commonly called “The ...
  3. [3]
    [PDF] The Isomorphism Theorems
    Theorem 14.1 (First Isomorphism Theorem). Let φ : V → W be a homomorphism between two vector spaces over a field F. (i) The kernel of φ is ...
  4. [4]
    3.2: Definitions of Homomorphisms and Isomorphisms
    Sep 19, 2021 · Intuitively, you can think of a homomorphism as a “structure-preserving” map: if you multiply and then apply homormorphism, you get the same ...
  5. [5]
    16.5: Ring Homomorphisms and Ideals - Mathematics LibreTexts
    Jun 4, 2022 · Table of contents. No headers. In the study of groups, a homomorphism is a map that preserves the operation of the group.
  6. [6]
    4.3: Image and Kernel - Mathematics LibreTexts
    Mar 5, 2022 · The kernel is the set of elements of G which map to the identity of H . The kernel is a subset of G , ...
  7. [7]
    10.1: Factor Groups and Normal Subgroups - Mathematics LibreTexts
    Sep 7, 2021 · If N is a normal subgroup of a group G , then the cosets of N in G form a group G ⁡ / N under the operation ( a ⁢ N ) ⁢ ( b ⁢ N ) = a ⁢ b ⁢ N .
  8. [8]
    [PDF] Lecture #14 of 24 ∼ October 19th, 2020 - Math 5111 (Algebra 1)
    Oct 19, 2020 · By using the first isomorphism theorem, we can construct isomorphisms of groups. In order to show that G/N is isomorphic to a group H, we search ...
  9. [9]
    [PDF] Algebra fact sheet An algebraic structure (such as group, ring, field ...
    Isomorphic algebraic structures are regarded as the same, and algebraic structures of each type are classified up to an isomorphism. ... Classification theorem of ...
  10. [10]
    [PDF] SUBGROUP SERIES I 1. Introduction If N is a nontrivial proper ...
    Roughly speaking, the Jordan–Hölder theorem says that a group determines its compo- sition series (assuming there is one at all). More precisely, a group ...
  11. [11]
    [PDF] NOTES FOR MATH 555, Spring 1999: FINITE GROUP THEORY
    Mar 30, 1999 · Now the Noether isomorphism theory, together with a simple induction argu- ment, yields the Jordan-Hölder Theorem. We state it only for ...<|separator|>
  12. [12]
    Arthur Cayley and the First Paper on Group Theory (Chapter 1)
    Introduction. Arthur Cayley's 1854 paper On the theory of groups, as depending on the symbolic equation θn = 1 inaugurated the abstract idea of a group [2].
  13. [13]
    (PDF) Arthur Cayley and the Abstract Group Concept - ResearchGate
    Aug 6, 2025 · 4. , On the theory of groups, as depending on the symbolic equation θn=1, Phil. Mag. 7(1854), 40–47,. reprinted in Collected Papers 125, vol ...
  14. [14]
    Felix Klein: The Erlangen Program | SpringerLink
    This book presents a historical account of Felix Klein's "Comparative Reflections on Recent Research in Geometry" (1872), better known as his "Erlangen Program.
  15. [15]
    The development of group theory - MacTutor
    Klein proposed the Erlangen Program in 1872 which was the group theoretic classification of geometry. Groups were certainly becoming centre stage in mathematics ...
  16. [16]
    [PDF] Heinrich Weber: Lehrbuch der Algebra (1895-96)
    More importantly, the idea that two isomorphic groups are in essence one and the same mathematical construct had been increasingly adopted. Weber's 1893 article ...
  17. [17]
    Theory of Groups of Finite Order
    Theory of Groups of Finite Order. Theory of Groups of Finite Order. Theory of ... Not only did Burnside prove many important theorems, but he also laid ...
  18. [18]
    Idealtheorie in Ringbereichen | Mathematische Annalen
    Noether, E. Idealtheorie in Ringbereichen. Math. Ann. 83, 24–66 (1921). https://doi.org/10.1007/BF01464225. Download citation. Received: 16 October 1920. Issue ...
  19. [19]
    Ideal Theory in Rings (Translation of "Idealtheorie in Ringbereichen ...
    Jan 11, 2014 · Abstract:This paper is a translation of the paper "Idealtheorie in Ringbereichen", written by Emmy Noether in 1920, from the original German ...
  20. [20]
    [PDF] Emmy Noether, Symmetry, and Women in Mathematics
    Idealtheorie in Ringbereichen Math. Annalen 83 (1921), 24-66. Ideal Theory in Rings. • Of fundamental importance in the development of modern algebra. • New ...<|separator|>
  21. [21]
    [PDF] Homological Algebra
    Algebraic Geometry. By SOLOMON LEFSCHETZ. 19. Homological Algebra. By HENRI CARTAN and SAMUEL EILENBERG. 20. The Convolution Transform. By I. I. HIRSCHMAN ...
  22. [22]
    [PDF] On the Structure of Abstract Algebras
    ON THE STRUCTURE OF ABSTRACT ALGEBRAS. BY GARRETT BIRKHOFF, Trinity College. [Communicated by MB P. HALL]. [Received 26 April, read. 3 June 1935].
  23. [23]
    B.L. Van Der Waerden, Moderne algebra, first edition (1930–1931)
    This chapter focuses on B.L. Van Derwaerden's book, Moderne Algebra, which was the first textbook of algebra systematically to present, on an abstract ...
  24. [24]
    [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.
  25. [25]
    [PDF] Module Fundamentals
    Note that by the first isomorphism theorem, every module is a quotient of a free module. Problems For Section 4.3. 1. Show that in Proposition 4.3.3, (2) can be ...
  26. [26]
    [PDF] Modules and Vector Spaces - Math@LSU
    (First isomorphism theorem) Let M and N be modules over the ring R and let f : M → N be an R-module homomorphism. Then. Page 8. 114. Chapter 3. Modules and ...
  27. [27]
    [PDF] some basic module-theoretic notions and examples
    Example 5.1. Let I = 2Z and J = 3Z in Z. Let ϕ: I → J by ϕ(x) = (3/2)x. This is an isomorphism of Z-modules. Even when two ideals are not principal, the ...
  28. [28]
    [PDF] MTH 302: Modules
    (Third Isomorphism theorem): Let M be an R-module and L,K < M such that. K ... [Dummit] Dummit and Foote, Abstract Algebra. [Lang] S. Lang, Algebra ...
  29. [29]
    [PDF] RES.18-011 (Fall 2021) Lecture 5: The Correspondence Theorem
    In particular, the integers modulo p, Zp, form a cyclic group of prime order; that is, any group of prime order p is isomorphic to Zp. Using Lagrange's Theorem ...
  30. [30]
    [PDF] the correspondence theorems for homomorphisms and
    The correspondence theorems describe objects associated to G/K in terms of objects associated to G and K, such as homomorphisms and subgroups.
  31. [31]
    [PDF] Important theorems about ring homomorphisms and ideals.
    Then ψ is an isomorphism from the quotient ring R/I to the ring ϕ(R). Hence, the rings R/I and ϕ(R) are isomorphic. 3. (The Correspondence Theorem.) Suppose ...
  32. [32]
    [PDF] atiyahmacdonald.pdf
    Chapter 2 Modules. Modules and module homomorphisms. Submodules and quotient modules. Operations on submodules. Direct sum and product. Finitely generated ...
  33. [33]
    [PDF] Supplement and Solution Manual for Introduction to Commutative ...
    May 29, 2021 · This note is based on my practice about Atiyah-MacDonald's book [3]. ... first isomorphism theorem. Also, This commutes with the map θk+1 : A ...
  34. [34]
    Correspondence theorem for rings. - Math Stack Exchange
    May 29, 2014 · Could someone provide a reference that includes a full and honest proof of the Correspondence Theorem for rings? Let A be a multiplicative ring ...Correspondence theorem in Ring theory - Math Stack ExchangeOne-to-one correspondence of ideals in the quotient also extends to ...More results from math.stackexchange.com
  35. [35]
    [PDF] Introduction to Commutative Algebra - OSU Math Department
    Apr 8, 2008 · with the middle line following from the Second Isomorphism Theorem. ... [1] M. F. Atiyah and I. G. Macdonald, Introduction to commutative algebra,.
  36. [36]
    [PDF] an introduction to the zariski topology - UChicago Math
    In this paper we will study the spectrum of a ring, which gives a way to define a topological space that can be created from any ring. This topological space, ...
  37. [37]
    Introduction to Modules and their Structure Theory
    As with groups and rings, we lead in Noether's isomorphism theorems for modules with a method to define maps from factor modules. Proposition 1.12. Given a map ...
  38. [38]
    [PDF] NOETHERIAN MODULES 1. Introduction In a finite-dimensional ...
    Theorem 2.8. For a Noetherian R-module M, each surjective R-linear map M → M is injective and thus is an isomorphism. Proof ...
  39. [39]
    [PDF] A Course in Universal Algebra
    Homomorphisms and the Homomorphism and Isomorphism Theorems . . . . ... Thus Birkhoff's theorem guarantees that for every Boolean algebra B there is an X and a.
  40. [40]
    Lemma 12.5.20 (05QB)—The Stacks project
    May 12, 2020 · If \beta , \delta are isomorphisms, \epsilon is injective, and \alpha is surjective then \gamma is an isomorphism. Proof. Immediate consequence ...
  41. [41]
    [PDF] An Introduction to Categories and Homological Algebra - IMJ-PRG
    Homological algebra is essentially linear algebra, no more over a field but over a ring and by extension, in abelian categories. When a field is replaced with a ...
  42. [42]
    [PDF] An Introduction to Homological Algebra
    Jun 1, 2011 · ... Weibel An introduction to homological algebra. 39 W. Bruns and J ... Lemma. 171. 6.4. Crossed Homomorphisms and H I. 174. 6.5. The Bar ...
  43. [43]
    [PDF] Foundations of Algebraic Topology
    By HASSLER WIIITNEY. Page 3. FOUNDATIONS. OF. ALGEBRAIC. TOPOLOGY. BY SAMUEL ElLENBERG. AND. NORMAN STEENROD ... lemma: LEMMA 4.3 (THE "FIVE" LEMMA). Let 11/I.l ...
  44. [44]
    13.6 Quotients of triangulated categories - Stacks Project
    Given a triangulated category and a triangulated subcategory we can construct another triangulated category by taking the “quotient”. The construction uses a ...
  45. [45]
    Correspondences and stable homotopy theory - Garkusha - 2023
    Sep 28, 2023 · By the proof of [14, Theorem 9.13] the latter are isomorphic in Mod ∼ S H ( k ) eff E $\widetilde{\operatorname{Mod}}_{SH(k)}^{\mathsf {eff}} ...<|control11|><|separator|>