Fact-checked by Grok 2 weeks ago

Inverse element

In abstract algebra, an for an element a in a set S equipped with an associative \cdot and a two-sided e is an element b \in S such that a \cdot b = e and b \cdot a = e. This two-sided inverse "reverses" the operation, restoring the identity, and is unique when it exists in such structures. In more general algebraic structures like monoids (sets with an associative operation and identity, but not necessarily inverses for all elements), an element may have a left inverse l where l \cdot a = e, or a right inverse r where a \cdot r = e, but these may not coincide or exist for every element. If an element has both a left and right inverse in a monoid, they are equal, forming a two-sided inverse. A key theorem states that if every element in a monoid has a left inverse, then the monoid is a group, where every element has a unique two-sided inverse. The is central to groups, defined as monoids where inverses exist for all ; here, the of a, denoted a^{-1}, satisfies (a^{-1})^{-1} = a and (a \cdot b)^{-1} = b^{-1} \cdot a^{-1}. Common examples include the additive inverse in the real numbers under addition, where the inverse of a is -a since a + (-a) = 0, the additive identity. Similarly, in the nonzero real numbers under multiplication, the inverse of a is $1/a because a \cdot (1/a) = 1, the multiplicative identity. Inverse elements enable solving equations like a \cdot x = e for x = a^{-1}, and they underpin structures in (e.g., rotations and reflections in symmetry groups) and (e.g., matrix inverses, where A^{-1} satisfies A \cdot A^{-1} = I, the identity matrix, provided \det(A) \neq 0). In non-associative structures like magmas, inverses may lack uniqueness or two-sidedness, but the idea extends to quasigroups where left and right inverses always exist.

Fundamental Definitions

Identity Element

In abstract algebra, an , also known as a element, is a concept in the of operations on sets. Given a set S equipped with a binary operation * : S \times S \to S, forming a magma, an element e \in S is a two-sided identity if it satisfies a * e = e * a = a for every a \in S. This property ensures that combining any element with the identity leaves the element unchanged, serving as a neutral point for the operation. Identity elements can be classified based on their behavior relative to the operation. A left identity satisfies e * a = a for all a \in S, while a right identity satisfies a * e = a for all a \in S. A two-sided identity combines both properties. In general magmas, left and right identities may exist independently and need not coincide, though in more structured settings like monoids, the identity is typically two-sided and unique. Common examples illustrate these concepts in familiar settings. Under addition, the integer 0 acts as the two-sided identity for the set of integers \mathbb{Z}, since a + 0 = 0 + a = a for any a \in \mathbb{Z}. Similarly, for multiplication on the nonzero real numbers \mathbb{R} \setminus \{0\}, the number 1 serves as the two-sided identity, as a \cdot 1 = 1 \cdot a = a for any a \in \mathbb{R} \setminus \{0\}. These cases highlight how identities preserve the structure of the operation without altering elements. The of a two-sided , when it exists, follows directly from the definition without requiring additional axioms like associativity. Suppose e and f are both two-sided identities in a magma (S, *). Then e = e * f (since f is a right identity) and e * f = f (since e is a left identity), so e = f. This proof establishes that no magma can have more than one two-sided identity, though finite sets or specific conditions (such as the operation being associative) can guarantee existence in certain cases. The of the emerged in the of during the , notably formalized by in her axiomatic treatment of in her seminal 1921 paper "Idealtheorie in Ringbereichen," influencing the modern standardization of algebraic structures that incorporate identities to facilitate concepts like inverses in subsequent sections. Noether's definition of rings did not require a multiplicative , though she specified for domains.

Basic Definition of Inverse

In abstract algebra, given a set S equipped with a * and an e \in S such that e * x = x * e = x for all x \in S, an element b \in S is called the inverse of an element a \in S if it satisfies a * b = e and b * a = e. This relation defines a two-sided inverse, where b acts as both a right inverse (satisfying a * b = e) and a left inverse (satisfying b * a = e) that coincide. The inverse of a, when it exists, is commonly denoted by a^{-1}, so that the defining equations become a * a^{-1} = a^{-1} * a = e. This notation emphasizes the "reciprocal" nature of the inverse with respect to the operation *, analogous to multiplicative inverses in arithmetic. The existence of an inverse for any element presupposes the presence of an identity element in the structure; without an identity, the concept of an inverse as defined here is not applicable. A simple example occurs in the set of integers \mathbb{Z} under the operation of addition, where the identity is $0 and the additive inverse of an integer a is -a, satisfying a + (-a) = (-a) + a = 0.

Left and Right Inverses

In algebraic structures equipped with an e, an element b is called a left inverse of an element a if b \cdot a = e. Similarly, an element c is a right inverse of a if a \cdot c = e. If both a left inverse and a right inverse exist for a and coincide (i.e., b = c), then b serves as a two-sided inverse of a, satisfying both b \cdot a = e and a \cdot b = e. In non-commutative settings, such as matrix multiplication, left and right inverses may exist independently without coinciding. For instance, consider the $2 \times 1 matrix A = \begin{pmatrix} 1 \\ 0 \end{pmatrix}. This has a left inverse B = \begin{pmatrix} 1 & 0 \end{pmatrix}, since B A = 1, but no right inverse exists, as any candidate C (a $1 \times 2 matrix) would yield A C of rank at most 1, unable to equal the $2 \times 2 identity matrix. In associative structures, the existence of both a left inverse b and a right inverse c for the same a implies that b = c, yielding a unique two-sided . This follows from the associativity condition: multiplying b \cdot a = e on the right by c gives b \cdot (a \cdot c) = e \cdot c, so b = b \cdot e = b \cdot (a \cdot c) = (b \cdot a) \cdot c = e \cdot c = c. Without associativity, however, a left inverse need not be a right inverse, even relative to the same identity. For example, consider the non-associative magma on the set \{1, a, b\} with identity $1 and multiplication table
\cdot1ab
11ab
aa1b
b11a
Here, b \cdot a = 1 (so b is a left inverse of a), but a \cdot b = b \neq 1 (so b is not a right inverse of a).

Properties in Basic Algebraic Structures

In Magmas

In a unital magma (M, \cdot, e), where e is the two-sided satisfying e \cdot x = x \cdot e = x for all x \in M, an element a \in M is said to have a left inverse b \in M if b \cdot a = e, a right inverse c \in M if a \cdot c = e, and a two-sided inverse d \in M if d \cdot a = a \cdot d = e. The concept of a magma as the foundational algebraic structure consisting of a set equipped with a binary operation was formalized in the 1940s by the Nicolas Bourbaki collective, emphasizing its role as the most general framework before imposing additional axioms like associativity or identity. The absence of associativity in magmas introduces significant challenges for inverses: the existence of a left inverse for an element does not guarantee a right inverse, and even when both exist, they need not coincide. This contrasts with associative structures, where a left inverse and right inverse for the same element would be equal via the relation b = b \cdot (a \cdot c) = (b \cdot a) \cdot c = c, but non-associativity invalidates such regrouping. For instance, consider a modified quaternion algebra over the reals where the standard multiplication table is altered only by setting i^2 = -1 + \epsilon j for a small nonzero real \epsilon, while keeping other products unchanged; here, the element i admits both a left inverse and a right inverse, but these inverses are distinct elements. A prominent example of a non-associative unital magma featuring two-sided inverses is the algebra of octonions \mathbb{O}, an 8-dimensional division algebra over the reals with basis \{1, e_1, \dots, e_7\} and a non-associative multiplication table derived from the Fano plane. Every nonzero octonion a \in \mathbb{O} has a unique two-sided inverse given by a^{-1} = \bar{a} / \|a\|^2, where \bar{a} denotes the conjugate (obtained by negating the imaginary parts) and \|a\| is the Euclidean norm satisfying \|a b\| = \|a\| \|b\| for all a, b \in \mathbb{O}. Despite this, the lack of associativity—evident in associators [a, b, c] = (a b) c - a (b c) \neq 0 for some triples—prevents \mathbb{O}^\times (the nonzero octonions under multiplication) from exhibiting group-like behavior, such as well-defined powers beyond alternativity. In general, the mere presence of two-sided inverses in a unital does not confer properties akin to those in groups, as non-associativity can cancellation, solvability of equations, and under repeated operations. For example, while (solving a x = b or x a = b uniquely for x when a \neq 0), the non-associative product complicates higher-order computations, limiting applications to contexts like exceptional Lie groups where alternativity suffices.

In Semigroups

In semigroups, which lack a required , the traditional notion of an inverse element is extended through the concept of regularity. An element a in a semigroup S is regular if there exists an element b \in S, called a weak inverse, such that aba = a. A stronger condition defines a generalized inverse b of a when both aba = a and bab = b hold. This generalized inverse acts analogously to an inverse relative to local identities within subsemigroups generated by a and b. aba = a, \quad bab = b A semigroup is termed regular, or von Neumann regular, if every element possesses at least one generalized inverse. This structure captures associative operations where elements can be "reversed" locally without a global unit, distinguishing it from more structured algebras like monoids. The terminology "von Neumann regular" draws from the parallel definition in ring theory, adapted to semigroups to emphasize self-inverse-like behavior under composition. Inverse semigroups form a key subclass of regular semigroups, where each element a has a unique generalized inverse b satisfying the above equations, and additionally, ab and ba are idempotents (i.e., (ab)^2 = ab and (ba)^2 = ba). This uniqueness ensures that the inverse behaves consistently across the semigroup, modeling partial symmetries akin to restrictions of group actions. Green's relations provide a framework for understanding how these inverses interact with the semigroup's structure, partitioning elements based on principal ideals. Specifically, the right relation \mathcal{R} equates a \mathcal{R} b if the principal right ideals generated by a and b coincide (Sa = Sb), and the left relation \mathcal{L} does so for left ideals (aS = bS). In a regular semigroup, a generalized inverse b of a satisfies a \mathcal{R} b and a \mathcal{L} b, placing them in the same \mathcal{H}-class (\mathcal{H} = \mathcal{R} \cap \mathcal{L}), where \mathcal{H}-classes often form groups containing the relevant idempotents. A representative example is the symmetric inverse semigroup on a set X, consisting of all partial bijections from X to X under . Each element, a partial bijection f, has a unique f^{-1}, also a partial bijection, satisfying the generalized inverse conditions and mapping domains and images appropriately, illustrating how inverses correspond to bijective transformations on restricted subsets.

In Monoids

In a monoid (M, \cdot, e), where \cdot is an associative binary operation and e is the , an element u \in M is called a unit if there exists an element v \in M such that u \cdot v = v \cdot u = e; here, v is the of u, denoted u^{-1}. The set of all units in M, denoted U(M), consists precisely of those elements possessing two-sided inverses. If an inverse exists for a unit u \in M, it is unique; specifically, if u \cdot v = v \cdot u = e and u \cdot w = w \cdot u = e, then v = w. Moreover, the product of two units is a unit: if u, v \in U(M), then u \cdot v \in U(M) with inverse given by (u \cdot v)^{-1} = v^{-1} \cdot u^{-1}. The set U(M) under the operation \cdot forms a group, known as the unit group of the monoid, inheriting associativity from M, with e as identity and inverses as defined. A representative example is the free generated by a set of letters under , where the is the empty word and the only is the empty word itself, as no non-empty word admits a two-sided . In , the of numbers (including 0) under has 0 but no non-zero s, since additive inverses do not exist within the set. Unlike groups, monoids do not require every element to be a , allowing for structures where invertibility is partial.

Inverses in Groups

Uniqueness of Inverses

In a group G with operation * and identity element e, if an element a \in G has an inverse b \in G such that a * b = e and b * a = e, then this inverse is unique. To prove this, suppose b and c are both two-sided inverses of a, so a * b = e = b * a and a * c = e = c * a. Multiplying the equation a * b = e on the left by c gives c * (a * b) = c * e, which by associativity equals (c * a) * b = c. Since c * a = e, this simplifies to e * b = c, or b = c. Thus, the inverse is unique relative to the identity e. This uniqueness extends to one-sided inverses in the sense that if both a left inverse and a right inverse exist for a, then they coincide and are the unique two-sided inverse. Specifically, if b * a = e (left inverse) and a * c = e (right inverse), then b = b * (a * c) = (b * a) * c = e * c = c. A concrete example occurs in the S_3, the group of all permutations of three elements under composition, which has order 6. Each element, such as the 3-cycle (1\ 2\ 3) whose inverse is (1\ 3\ 2), or the transposition (1\ 2) which is its own inverse, has a unique inverse that "undoes" the permutation to yield the . As a key consequence, the uniqueness of inverses allows solving equations in groups: for a * x = b, multiplying on the left by a^{-1} yields the unique solution x = a^{-1} * b. This property was established in the foundational development of group theory by Camille Jordan in his 1870 treatise Traité des substitutions et des équations algébriques, where he systematically treated permutation groups and their structural properties, including inverses.

Group Operations with Inverses

In groups, the existence of inverses facilitates the solution of equations of the form a * x = b, where a, b \in G and G is the group under operation *. By left-multiplying both sides by a^{-1}, one obtains x = a^{-1} * b, leveraging the uniqueness of inverses to ensure a unique solution. Similarly, for equations x * a = b, right-multiplication by a^{-1} yields x = b * a^{-1}. These manipulations rely on the associative property and the inverse axiom, enabling cancellation laws that distinguish groups from weaker structures. The presence of inverses extends the definition of powers to negative exponents, defining a^{-n} = (a^{-1})^n for positive integers n and a \in G. This follows from the relation (a^n)^{-1} = a^{-n}, proved by induction: for n=1, it holds by definition, and assuming it for k, then a^{-(k+1)} = (a^{-k}) * a^{-1} = (a^{k+1})^{-1}. Consequently, the exponent laws a^{m+n} = a^m * a^n and (a^m)^n = a^{mn} hold for all integers m, n, allowing full integer exponentiation within the group. Subgroups, as subsets closed under the group operation and containing the identity, must also be inverse-closed: if h \in H \subseteq G, then h^{-1} \in H to preserve the subgroup axioms. This closure ensures that every subgroup inherits the inverse property, forming a group under the restricted operation. For instance, in the cyclic group \mathbb{Z}_n under addition modulo n, the inverse of k is n - k \pmod{n} (or -k \pmod{n}), since k + (n - k) \equiv 0 \pmod{n}, where 0 is the identity./03:_Groups/3.09:_Subgroups) In any group, every element a generates a cyclic subgroup \langle a \rangle = \{a^k \mid k \in \mathbb{Z}\}, which incorporates inverses via negative exponents to form a complete cyclic structure. This property underscores that groups are precisely those algebraic structures where inverses enable every element to produce such invertible cyclic subgroups. In applications, such as symmetry groups in physics, inverses correspond to "undoing" transformations; for example, in the rotation group SO(3), the inverse of a rotation by angle \theta around an axis is rotation by -\theta, preserving physical symmetries like those of particles or crystals.

In Rings

In a ring R, the addition operation forms an , so for every element r \in R, there exists a unique -r \in R such that r + (-r) = 0_R, where $0_R is the ; this follows directly from the ring axioms requiring the additive structure to be an . Multiplicative inverses in a ring, when they exist, are defined with respect to the multiplicative identity $1_R (assuming the ring has unity, as is standard in much of ); an element u \in R is called a unit if there exists u^{-1} \in R such that u \cdot u^{-1} = u^{-1} \cdot u = 1_R. The set of all units in R, denoted R^\times, forms a group under the ring's multiplication operation, known as the multiplicative group of units. For example, in the ring of integers \mathbb{Z}, the units are precisely \pm 1, as these are the only elements whose products yield the multiplicative identity 1. In the ring of n \times n matrices over a field F, denoted M_n(F), the units consist of the invertible matrices, which are exactly those with nonzero determinant (since the determinant must be a unit in F, and the nonzero elements of F are the units). In general rings, multiplicative inverses exist only for units and not necessarily for all nonzero elements, distinguishing rings from fields where every nonzero element is a unit.

In Fields

In a field F, every nonzero element has a unique multiplicative inverse, and every element has an additive inverse, making F an abelian group under addition and the nonzero elements forming an abelian group under multiplication. This universal invertibility for nonzero elements enables division by any such element within the field, distinguishing fields from rings where units are only a subset of the elements. A fundamental example is the field of rational numbers \mathbb{Q}, constructed as fractions of integers where, for any nonzero r = \frac{p}{q} with integers p \neq 0 and q \neq 0, the multiplicative inverse is \frac{q}{p}. This structure ensures that \mathbb{Q} satisfies the field axioms, allowing operations like solving linear equations through division without leaving the set. The multiplicative inverse satisfies the equation a \cdot a^{-1} = 1 for all a \in F with a \neq 0, where $1 is the multiplicative identity. In finite fields, such as \mathbb{F}_p = \mathbb{Z}/p\mathbb{Z} for prime p, Fermat's Little Theorem provides a method to compute inverses: the inverse of nonzero a is a^{p-2} \mod p, since a^{p-1} \equiv 1 \mod p implies a \cdot a^{p-2} \equiv 1 \mod p. Fields play a central role in as commutative domains where every nonzero element is a , enabling the construction of spaces, extensions, and solutions to equations. This property underpins applications in and , where the absence of zero divisors combined with full invertibility ensures unique factorization and division algorithms.

Applications to Matrices and Linear Algebra

Standard Matrix Inverses

In linear algebra over a field F, a square matrix A \in M_n(F) is said to be invertible if there exists another square matrix B \in M_n(F) such that AB = BA = I_n, where I_n is the n \times n identity matrix. This matrix B, when it exists, is unique and denoted A^{-1}. A square matrix A is invertible if and only if its determinant is nonzero, i.e., \det(A) \neq 0. This condition ensures that A has full rank and is thus nonsingular. One explicit method to compute the inverse uses the adjugate matrix: A^{-1} = \frac{1}{\det(A)} \adj(A), where \adj(A) is the adjugate, the transpose of the cofactor matrix of A. This formula holds provided \det(A) \neq 0, and it derives from Cramer's rule applied to the system defining the inverse. For a $2 \times 2 matrix A = \begin{pmatrix} a & b \\ c & d \end{pmatrix} with \det(A) = ad - bc \neq 0, the inverse is given by A^{-1} = \frac{1}{ad - bc} \begin{pmatrix} d & -b \\ -c & a \end{pmatrix}. This special case follows directly from the adjugate formula and is computationally efficient for small dimensions. Invertible matrices satisfy several key properties under multiplication. If A and B are invertible, then their product AB is invertible with (AB)^{-1} = B^{-1} A^{-1}. More generally, the set of invertible n \times n matrices over F forms the general linear group \mathrm{GL}_n(F), which is the group of units in the matrix ring M_n(F). A primary application of matrix inverses is solving linear systems. For an invertible matrix A and vector b \in F^n, the system A x = b has the unique solution x = A^{-1} b. This approach transforms the problem into a matrix-vector multiplication, though Gaussian elimination is often preferred for numerical stability in practice.

Generalized Matrix Inverses

The Moore-Penrose pseudoinverse, denoted A^+ for an m \times n A, provides a of the to arbitrary real or matrices, including those that are singular or rectangular. It is uniquely defined as the matrix satisfying the four Penrose equations: A A^+ A = A, A^+ A A^+ = A^+, (A A^+)^H = A A^+, and (A^+ A)^H = A^+ A, where ^H denotes the conjugate transpose (or transpose for real matrices). These conditions ensure that A^+ acts as a partial , projecting onto the range of A while minimizing the Euclidean norm of the solution in least-squares problems. Generalized inverses encompass several types beyond the full Moore-Penrose version, including left and right pseudoinverses tailored to matrices with specific properties. A left pseudoinverse G satisfies A G A = A and G A G = G, existing for matrices with full column , while a right pseudoinverse satisfies the same but applies to full row cases. The Moore-Penrose pseudoinverse subsumes these as a "full" or complete generalized inverse by additionally enforcing the symmetry conditions, making it applicable to all matrices. Computationally, the pseudoinverse is efficiently obtained via the singular value decomposition (SVD) of A = U \Sigma V^H, where A^+ = V \Sigma^+ U^H and \Sigma^+ reciprocates the nonzero singular values while setting zero singular values to zero. For a rank-deficient , such as A = \begin{pmatrix} 1 & 1 \\ 1 & 1 \end{pmatrix}, the pseudoinverse A^+ = \frac{1}{4} \begin{pmatrix} 1 & 1 \\ 1 & 1 \end{pmatrix} yields the minimum-norm least-squares to A x = b, minimizing \|A x - b\|_2 and then \|x\|_2 among minimizers. Key properties include existence for every matrix over the reals or complexes, uniqueness of the Moore-Penrose form, and the fact that it reduces to the standard inverse when A is invertible. Historically, the concept originated with E. H. Moore's work on general analysis in 1920 and was independently formalized by Roger Penrose in 1955, with applications in statistics for regression analysis and in control theory for system identification and stabilization.

Inverses for Functions and Morphisms

Functional Inverses

In mathematics, a function f: X \to Y between sets X and Y has an inverse function g: Y \to X if the composition f \circ g equals the identity function on Y (i.e., f(g(y)) = y for all y \in Y) and g \circ f equals the identity function on X (i.e., g(f(x)) = x for all x \in X). This condition ensures that g precisely undoes the action of f, and vice versa. The , when it exists, is denoted by f^{-1}, so f^{-1}: Y \to X satisfies f \circ f^{-1} = \mathrm{id}_Y and f^{-1} \circ f = \mathrm{id}_X. If an exists for f, it is unique; supposing another h: Y \to X also satisfies the inverse conditions, then h = f^{-1} follows directly from composing with f or f^{-1}. Additionally, the operation reverses under composition: (f^{-1})^{-1} = f, meaning applying the twice returns . A concrete example is the function f: \mathbb{R} \to \mathbb{R} defined by f(x) = 2x, which has inverse f^{-1}(x) = x/2; verifying, f(f^{-1}(x)) = 2(x/2) = x and f^{-1}(f(x)) = (2x)/2 = x for all real x. In , a function f: X \to Y admits an inverse if and only if it is bijective, meaning both injective (one-to-one) and surjective (onto); bijectivity guarantees the existence and uniqueness of f^{-1}. Equivalently, f is invertible precisely when it establishes a one-to-one correspondence between X and Y. For non-bijective functions, partial inverses may still be defined on restricted domains or codomains. If f: X \to Y is injective but not surjective, a left inverse g: Y \to X exists such that g \circ f = \mathrm{id}_X, though g is defined only partially on the image of f. Conversely, if f is surjective but not injective, a right inverse h: Y \to X satisfies f \circ h = \mathrm{id}_Y, again partial in the sense that it selects one preimage per element in Y. These notions appear, for instance, in defining branches of inverse trigonometric functions like \arcsin, which inverts the sine function restricted to [-\pi/2, \pi/2].

Inverses in Category Theory

In category theory, the notion of an inverse generalizes the of an from the to arbitrary . A f: A \to B in a \mathcal{C} is said to have an inverse if there exists a g: B \to A such that f \circ g = \mathrm{id}_B and g \circ f = \mathrm{id}_A, where \mathrm{id} denotes the identity . This definition captures the idea of invertibility in a structure-preserving manner, analogous to bijective functions between sets serving as inverses. Such invertible morphisms are precisely the isomorphisms in the category. An isomorphism between objects A and B is a morphism f: A \to B equipped with an inverse g: B \to A satisfying the above compositions. In the Hom-set \mathrm{Hom}_\mathcal{C}(A, B), the set of all morphisms from A to B, the isomorphisms form a group under composition whenever \mathrm{Hom}_\mathcal{C}(A, A) is considered, with the identity as the unit element. A category in which every morphism is an isomorphism is called a groupoid, highlighting how inverses permeate the entire structure. While strict inverses define isomorphisms, a weaker notion arises with equivalences of categories, which involve functors that are inverses up to natural isomorphism. Specifically, two categories \mathcal{C} and \mathcal{D} are equivalent if there exist functors F: \mathcal{C} \to \mathcal{D} and G: \mathcal{D} \to \mathcal{C} such that there are natural isomorphisms \eta: \mathrm{id}_\mathcal{C} \to G \circ F and \epsilon: F \circ G \to \mathrm{id}_\mathcal{D}. This provides a sense of invertibility that preserves the essential structure without requiring exact equality of morphisms. For a , consider the \mathbf{Grp} of groups and group homomorphisms. Here, the isomorphisms are precisely the bijective group homomorphisms, each of which admits an that is also a group homomorphism. In more advanced settings, such as monoidal categories, the object \mathbf{1} is invertible, meaning there exists an object I (often \mathbf{1} itself) and isomorphisms \lambda: I \otimes \mathbf{1} \to I and \rho: \mathbf{1} \otimes I \to I satisfying coherence conditions, emphasizing the role of inverses in tensor structures. More generally, an object X in a monoidal category is invertible if there exists Y such that X \otimes Y \cong \mathbf{1} \cong Y \otimes X, allowing X to act as a unit up to isomorphism.

Advanced Generalizations

In Semirings

A semiring is an algebraic structure generalizing a ring by forgoing the requirement of additive inverses for its elements. Formally, it consists of a nonempty set S equipped with two binary operations + (addition) and \cdot (multiplication), where (S, +) is a commutative monoid with identity $0, (S, \cdot) is a monoid (typically with identity $1), multiplication distributes over addition on both sides, and $0 is absorbing under multiplication ($0 \cdot a = a \cdot 0 = 0). Multiplicative units exist as the identity $1 satisfying $1 \cdot a = a \cdot 1 = a for all a \in S, but additive inverses generally do not, so there is no -a \in S such that a + (-a) = 0 for arbitrary a \neq 0. For instance, the semiring of nonnegative integers \mathbb{N}_0 = \{0, 1, 2, \dots\} under standard addition and multiplication exemplifies this, as only $0 has an additive inverse. The Boolean semiring \mathcal{B} = (\{0,1\}, \lor, \land, 0, 1), where \lor denotes (addition) and \land denotes (multiplication), further illustrates the absence of additive inverses while preserving a multiplicative . Here, $1 acts as the since a \land 1 = a for a \in \{0,1\}, but no element b satisfies $1 \lor b = 0, precluding additive inverses. Rings represent a special case of s where additive inverses do exist, enabling subtraction. In idempotent semirings, where addition satisfies a + a = a for all a, the Kleene star operation a^* generalizes inverses by solving iterative equations. A Kleene algebra extends an idempotent semiring with a unary star operator satisfying axioms such as a^* = 1 + a \cdot a^* and (a \cdot b)^* = a \cdot (b \cdot a)^* \cdot b, analogous to the inverse of (I - a) in matrix semirings for regular languages. This allows "inversion" in the sense of closure under summation of powers, as in x = 1 + a x yielding x = a^*. In the tropical (min-plus) semiring \mathbb{T} = (\mathbb{R} \cup \{\infty\}, \min, +, \infty, 0), addition is the minimum operation and multiplication is standard , with \infty as the additive identity and $0 as the multiplicative identity. Additive inverses are absent, as no b satisfies \min(a, b) = \infty for finite a. However, the negation map \phi: x \mapsto -x (extended appropriately with \phi(\infty) = -\infty) induces an isomorphism to the dual max-plus semiring (\mathbb{R} \cup \{-\infty\}, \max, +, -\infty, 0), providing a structural "inverse" transformation between these semirings rather than element-wise inverses. Semirings often carry a partial order a \leq b a + b = b, particularly in idempotent cases, which turns the additive structure into a join-semilattice. This enables operations as generalized inverses for inequalities: given order-preserving multiplication, the right b / a is the greatest x such that a \cdot x \leq b, and similarly for the left . These residuals facilitate solving a \cdot x \leq b in ordered settings, akin to division in ordered groups. In computer science, semirings model formal languages via formal power series, where addition corresponds to union and to concatenation; this framework supports weighted automata and Kleene's theorem for recognizing languages, with applications in parsing and optimization.

Quasi-Inverses and U-Semigroups

In semigroup theory, a quasi-inverse of an element a in a semigroup (S, \cdot) is an element b \in S such that a \cdot b \cdot a = a. This condition provides a weaker notion of invertibility compared to groups, where an inverse b satisfies a \cdot b = b \cdot a = e for an identity e, as no identity or symmetric condition is required here. Quasi-inverses play a key role in defining regular semigroups, which are those where every element has at least one quasi-inverse, allowing the study of elements with "partial reversibility" without full symmetry. U-semigroups are semigroups in which every possesses a quasi-inverse. In a U-semigroup, if b is the quasi-inverse of a (satisfying a \cdot b \cdot a = a), then b also satisfies b \cdot a \cdot b = b, establishing b as a full inverse and rendering the structure an inverse semigroup. This equivalence highlights how uniqueness of quasi-inverses enforces the additional symmetry needed for inverse semigroups, where each has precisely one full inverse obeying both equations. These properties relate to regularity, as U-semigroups are a special class of regular semigroups with this uniqueness constraint. The full transformation semigroup T_X on a nonempty set X, consisting of all functions from X to X under composition, illustrates quasi-inverses through partial bijections. For a transformation f \in T_X, a quasi-inverse g is a partial bijection such that f \circ g \circ f = f, effectively "reversing" f on its image. The subsemigroup of all partial bijections on X forms the symmetric inverse semigroup I_X, where elements admit unique full inverses, demonstrating how quasi-inverses in broader transformation semigroups connect to inverse structures. The concepts of quasi-inverses and U-semigroups emerged within developments in semigroup theory, building on earlier work in regularity and inverse structures by researchers including J. M. Howie, whose foundational texts systematized these extensions.

Galois Connections as Inverses

A Galois connection provides a duality between partially ordered sets (posets) that mimics the of inverses in a generalized, order-preserving manner. Given two posets (P, \leq_P) and (Q, \leq_Q), a Galois connection consists of two monotone (order-preserving) functions f: P \to Q and g: Q \to P satisfying the condition that for all x \in P and y \in Q, f(x) \leq_Q y \quad \iff \quad x \leq_P g(y). This equivalence establishes f as the lower adjoint and g as the upper adjoint to each other, capturing a bidirectional relationship without requiring bijectivity. The inverse-like of a Galois connection arises because f and g as approximate inverses, particularly on their fixed points. Specifically, the compositions gf: P \to P and fg: Q \to Q are monotone idempotent : gf forms a on P (extensive and idempotent), while fg forms an interior (contractive and idempotent) on Q. The fixed points of gf (elements x \in P where gf(x) = x) and the fixed points of fg (elements y \in Q where fg(y) = y) are isomorphic via f and g, providing a precise duality where the maps behave as true inverses between these subposets of closed and open elements, respectively. Key properties follow directly from the adjointness condition, including the closure equations fgf = f, \quad gfg = g, which confirm the idempotence and confirm that f and g are inverses up to the respective operators. In the antitone variant of Galois connections—where f and g are order-reversing—the duality strengthens to provide full inverses on the entire posets under certain completeness assumptions, such as when the posets are complete lattices, leading to a contravariant equivalence. A representative example occurs in power set lattices, where the posets are the power sets \mathcal{P}(X) and \mathcal{P}(Y) of sets X and Y, ordered by subset inclusion \subseteq. For a binary relation R \subseteq X \times Y, define f(A) = \{y \in Y \mid \forall x \in A, \, x \, R \, y\} (the set of common successors) and g(B) = \{x \in X \mid \forall y \in B, \, x \, R \, y\} (the set of common predecessors); adjusting for monotonicity via upset or downset constructions yields a Galois connection whose adjoint condition directly encodes subset inclusion relations between images. This setup illustrates how Galois connections generalize subset inclusions into dual operations on collections of subsets. Galois connections find applications in , where they model the duality between syntactic entailment (ordered by ) and semantic (ordered by model ), enabling approximations of provability via operators. In , they underpin , such as functional and multivalued dependencies, by deriving closed sets of attributes from schemas to optimize query and data mining tasks like .

References

  1. [1]
    AATA Definitions and Examples - Abstract Algebra
    ... inverse element in G, denoted by , a − 1 , such that ... Sophus Lie, a contemporary of Klein, used group theory to study solutions of partial differential ...
  2. [2]
    [PDF] Left and right inverses
    If S is a monoid and a has a left inverse, then a has at most one right inverse. Moreover, if it has a right inverse, then it is equal to the left inverse.Missing: abstract | Show results with:abstract
  3. [3]
    [PDF] Algebra.pdf - UChicago Math
    A group is a monoid in which each element m has an inverse element m−1, so that mm−1 = e = m−1m. A homomorphism f : M −→ N of monoids is a function f ...
  4. [4]
    [PDF] Definition Of Inverse In Math
    - Inverse element: For a given element, its inverse is another element that combines with it to produce the identity element. An example in addition: - 5 + (-5) ...
  5. [5]
    6.3 - The Inverse of a Square Matrix
    The inverse of a square matrix A, denoted by A -1 , is the matrix so that the product of A and A -1 is the Identity matrix.
  6. [6]
    [PDF] Basic notions of abstract algebra A magma is a set {g 1,g2 ...
    There is no notion of inverse element with respect to the multiplication. A field K is a set together with two composition laws + and · (addition and.
  7. [7]
    [PDF] binary opera- tions, magmas, monoids, g - UMD MATH
    The magma (Z, +) is associative and has 0 as its identity element. The magma (N, +) is also associative with 0 as its identity element. If n > 0, then the magma.
  8. [8]
    AlgebraicStructures - Computer Science
    If there is an identity element, it is easy to prove that it is unique: Let e and e' be identities. Then e*e' = e' (by the e*x=x rule) and e*e' = e (by the x* ...
  9. [9]
    [PDF] 4 Proofs in group theory
    The fact that each element has a unique inverse means that the identity e must occur exactly once in each row and each column. However, there is a slightly ...
  10. [10]
    13.3 Identity
    An identity element with respect to a binary operation is an element such that when a binary operation is performed on it and any other given element, the ...Missing: algebra | Show results with:algebra
  11. [11]
    [PDF] Why all rings should have a 1 - MIT Mathematics
    Jun 15, 2018 · Should the definition of ring require the existence of a multiplicative iden- tity 1? Emmy Noether, when giving the modern axiomatic definition ...
  12. [12]
    Ideal Theory in Rings (Translation of "Idealtheorie in Ringbereichen ...
    Jan 11, 2014 · This paper is a translation of the paper "Idealtheorie in Ringbereichen", written by Emmy Noether in 1920, from the original German into English.
  13. [13]
    [PDF] Notes on Abstract Algebra
    (3) =⇒ (2): Given α, we assume there is a left inverse β as in (3). If γ ... (3) =⇒ (2): Given α, we assume there is a right inverse β as in (3). If ...
  14. [14]
    [PDF] Definitions, Theorems and Exercises Math 332 Abstract Algebra ...
    Apr 20, 2017 · If < has an identity element, the identity element is unique. Definition 1.1.7. Let A be a set, and let < be a binary operation of A. Let e ...
  15. [15]
    [PDF] The Ring of Integers
    (d) Every element has an additive inverse: If n ∈ Z, there is an element −n ∈ Z such that n + (−n) = 0 and (−n) + n = 0. (e) Addition is commutative: If m, n ∈ ...
  16. [16]
    [PDF] Section 3: Fundamental Theorems about Groups
    i.e., an element of the semigroup has a “left inverse” but not a “right inverse” (or vice versa). If an element e satisfies e∗x = x for all x in S, we'll ...
  17. [17]
    [PDF] Left and right inverses; pseudoinverse - MIT OpenCourseWare
    = (AT A)−1 AT is a left inverse of A. (There may be other left in verses as ... or its transpose has a nonzero nullspace. Right inverse. If A has full ...Missing: abstract | Show results with:abstract
  18. [18]
    [PDF] 4. Matrix inverses
    Left inverse: 𝑋 is a left inverse of 𝐴 if. 𝑋𝐴 = 𝐼. 𝐴 is left-invertible if it has at least one left inverse. Right inverse: 𝑋 is a right inverse of 𝐴 if.Missing: abstract | Show results with:abstract
  19. [19]
    specific magma examples - abstract algebra - Math Stack Exchange
    Sep 4, 2014 · an example of a magma with an identity such that there is an element with exactly 2 left inverses but only one right inverse. For the first ...A ring element with a left inverse but no right inverse?Small magmas with identity, inverses, and almost associativityMore results from math.stackexchange.com
  20. [20]
    magma in nLab
    Aug 21, 2024 · A magma (or binary algebraic structure, or, alternatively, a mono-binary algebra) ( S , ⋅ ) (S,\cdot) is a set equipped with a binary operation on it.
  21. [21]
    [PDF] Chapter 7 Nicolas Bourbaki: Theory of Structures
    The influence of Bourbaki, both as a group and through the works of its individual members, thus played a central role in shaping the development of algebraic ...
  22. [22]
    [PDF] The Octonions - SISSA People Personal Home Pages
    Dec 21, 2001 · The element i then has both right and left inverses, but they are not equal. (We thank David Rusin for this example.) There are three levels of ...
  23. [23]
    [PDF] Binary Operations, Monoids, and Groups - CSUSM
    We sometimes call invertible element in a monoid units. We can rephrase ... If a ∈ A has an inverse b in the monoid A, then b is the inverse of a in M as well.
  24. [24]
    [PDF] Introduction to Abstract Algebra (Math 113)
    For a ∈ G there is only one element which behaves like the inverse of a. Proof. Assume a ∈ G has 2 inverses, b, c ∈ G. Then: (a ∗ b) = e c ∗ (a ∗ b) ...
  25. [25]
    [PDF] adjoining universal inverses to families of elements of free monoids
    Aug 29, 2025 · Abstract. Let X be the free monoid on a generating set X, and suppose one adjoins to X universal. 2-sided inverses to a finite set S of its ...
  26. [26]
    [PDF] Math 403 Chapter 5 Permutation Groups: 1. Introduction
    (a) Definition: The symmetric group Sn is the group of all permutations of the set 11, 2, ..., nl. Example: The group S3 consists of six elements.
  27. [27]
    Traité des substitutions et des équations algébriques - Internet Archive
    Dec 3, 2008 · Traité des substitutions et des équations algébriques. by: Jordan, Camille, 1838-1922. Publication date: 1870. Topics: Galois theory, Groups, ...Missing: inverses | Show results with:inverses
  28. [28]
    Manipulating equations in groups - Groupprops
    Mar 31, 2022 · This article is about the basic rules of manipulating equations involving groups. It is written mainly for beginners to the subject of group theory.2The rules for manipulating... · 2.4Not every equation can be... · 3.2What's not true<|control11|><|separator|>
  29. [29]
    [PDF] Properties of Powers in Groups
    We proved that inverses of elements in groups are unique, and we know gg-1 = g-1g = e, by definition of inverse. Thus, by uniqueness, we must have h = g, so (g- ...
  30. [30]
    [PDF] Cyclic Group Supplement Theorem 1. Let 𝑔 be an element of a ...
    Also, Z = ⟨−1⟩ because 𝑘 = (−𝑘) ⋅ (−1) for each 𝑘 ∈ Z. 7. Z𝑛 is a cyclic group under addition with generator 1. Theorem 4. Let 𝑔 be an element of a group 𝐺.
  31. [31]
    [PDF] 7 Symmetry and Group Theory - Penn Math
    Notice also that rotating by 0 serves as the identity element, and that each of the rotations have an inverse. Finally, rotations in space are always ...
  32. [32]
    16.1: Rings, Basic Definitions and Concepts - Mathematics LibreTexts
    Aug 16, 2021 · By Theorem 11.3.3, the multiplicative inverse of a ring element is unique, if it exists. For this reason, we can use the notation \(u^{-1}\) ...Basic Definitions · Multiplicative Inverses in Rings · Universal Concepts...
  33. [33]
    Units in a Ring - Socratica
    Rings have two operations: + + + and × \times ×. Any ring R R R is a commutative group under addition, so every element has an additive inverse.
  34. [34]
    16.3: Rings - Mathematics LibreTexts
    Jun 4, 2022 · A division ring is a ring with an identity, in which every nonzero element in is a unit ; that is, for each a ∈ R with a ≠ 0 , there exists a ...<|separator|>
  35. [35]
    [PDF] Chapter 4 - Matrices and Matrix Rings
    The highlight of the chapter is the theorem that a square matrix is a unit in the matrix ring iff its determinant is a unit in the ring. This chapter concludes ...
  36. [36]
    16.2: Fields - Mathematics LibreTexts
    Aug 16, 2021 · A field is a commutative ring with unity such that each nonzero element has a multiplicative inverse. In this chapter, we denote a field ...
  37. [37]
    7.4: D- Basic Terminology and Results from Algebra
    Sep 5, 2021 · If the multiplication also defines a group on the set of nonzero elements, then the ring is a field.
  38. [38]
    16.4: Integral Domains and Fields - Mathematics LibreTexts
    Jun 4, 2022 · Therefore, units of this ring are ± 1 and ± i ; hence, the Gaussian integers are not a field. We will leave it as an exercise to prove that the ...
  39. [39]
    18.1: Fields of Fractions - Mathematics LibreTexts
    Jun 4, 2022 · The rational numbers are certainly a field. In fact, it can be shown that the rationals are the smallest field that contains the integers.
  40. [40]
    2.1: Fields - Mathematics LibreTexts
    Sep 14, 2021 · (Multiplicative inverse) For all a , b , c ∈ F , a ⋅ ( b + c ) = a ⋅ b + a ⋅ c . (Distributive property I) For all a , b , c ∈ F , ( a + b ) ⋅ ...
  41. [41]
    Divisibility - American Mathematical Society
    , we have p ∤ a. By Fermat's Little Theorem (Theorem 7.5), a · ap−2 ≡ 1. Hence, [ap−2] is the multiplicative inverse of [a], so Zp is a field. D.
  42. [42]
    The Inverse of a Matrix — Linear Algebra, Geometry, and Computation
    Definition. A matrix A is called invertible if there exists a matrix C such that. AC=I and CA=I. In that case C is called the inverse of A.
  43. [43]
    [PDF] 2.5 Inverse Matrices - MIT Mathematics
    The matrix A−1 is called “A inverse.” DEFINITION The matrix A is invertible if there exists a matrix A−1 that “inverts” A : Two-sided inverse. A−1A = I ...
  44. [44]
    [PDF] Invertibility and Properties of Determinants - Sites at Lafayette
    A square matrix A is invertible if and only if detA 6= 0. In a sense, the theorem says that matrices with determinant 0 act like the number 0–they don't.
  45. [45]
    Math 21b: Determinants
    The determinant of a square matrix A detects whether A is invertible: If det(A)=0 then A is not invertible (equivalently, the rows of A are linearly dependent.Missing: iff | Show results with:iff
  46. [46]
    [PDF] Determinant and the Adjugate
    The formulae presented in these notes for the determinant and the inverse of a matrix are mainly of theoretical interest. They often can be used in proofs ...
  47. [47]
    [PDF] 3.2 Determinants and Matrix Inverses
    We are now able to define the adjugate of an arbitrary square matrix and to show that this formula for the inverse remains valid (when the inverse exists).
  48. [48]
    [PDF] The inverse of a 2 × 2 matrix - Mathcentre
    A simple formula for the inverse. In the case of a 2 × 2 matrix A = a b. c d ! a simple formula exists to find its inverse: if. A = a b. c d ! then. A−1 = 1 ad ...
  49. [49]
    Matrix Inverses
    The inverse of a product of several invertible matrices is the product of the inverses, in the opposite order; the proof is the same.
  50. [50]
    [PDF] arXiv:1110.6882v1 [math-ph] 31 Oct 2011
    Oct 31, 2011 · Moore-Penrose pseudoinverse. The notion of Moore-Penrose pseudoinverse was introduced by E. H. Moore [1] in 1920 and rediscovered by R.Missing: original | Show results with:original
  51. [51]
    Inverse Functions
    An inverse function reverses the operation done by a particular function. In other words, whatever a function does, the inverse function undoes it. In this ...
  52. [52]
    Inverse function definition - Math Insight
    An inverse function is a function that undoes the action of the another function. A function g is the inverse of a function f if whenever y=f(x) then x=g(y). ...
  53. [53]
    Calculus I - Inverse Functions - Pauls Online Math Notes
    Nov 16, 2022 · In this section we will define an inverse function and the notation used for inverse functions. We will also discuss the process for finding ...
  54. [54]
    [PDF] Inverse Functions
    A function g: B → A is called an inverse function for f if it satisfies the following condition: For every a ∈ A and b ∈ B, f(a) = b if and only if g(b) = a. ...
  55. [55]
    Functions:Inverses - Department of Mathematics at UTSA
    Nov 6, 2021 · In mathematics, an inverse function (or anti-function) is a function that "reverses" another function.
  56. [56]
    [PDF] Review of Set Theory - UCSB Math
    Mar 30, 2009 · A function f : A → B is invertible if and only if it is a bijection. Furthermore, an invertible function has a unique inverse function. Proof.
  57. [57]
    4.6 Bijections and Inverse Functions
    No matter what function f we are given, the induced set function f−1 is defined, but the inverse function f−1 is defined only if f is bijective.
  58. [58]
    [PDF] Partial Inverse Trigonometric Functions
    These partial inverses are called arcsine, arccosine, and arctangent. We'll begin this chapter with a review of the definition of the most famous example of a ...
  59. [59]
    inverse in nLab
    Sep 16, 2023 · An inverse of a morphism f : X → Y f \colon X \to Y in a category or unital magmoid (or an element of a monoid or unital magma) is another morphism.
  60. [60]
    isomorphism in nLab
    Dec 12, 2024 · An isomorphism is an invertible morphism, hence a morphism with an inverse morphism. Two objects of a category are said to be isomorphic if there exists an ...
  61. [61]
    groupoid in nLab
    Oct 26, 2025 · As such this is the sub-2-category of Cat on those (small) categories all whose morphisms are invertible morphism.
  62. [62]
    Grp in nLab
    Jul 7, 2025 · 1. Definition. Grp is the category with groups as objects and group homomorphisms as morphisms. Similarly there is the full subcategory FinGrp ...
  63. [63]
    semiring in nLab
    Apr 11, 2025 · However, more recently, rings have been defined with multiplicative units, and thus nowadays the primary definition of semiring used is the ...
  64. [64]
    Semirings :: Jens Zumbrägel
    The Boolean semiring is a semifield, i.e. all nonzero elements have a multiplicative inverse. Interestingly, it is the only finite proper semifield. Here is ...
  65. [65]
    [PDF] A Completeness Theorem for Kleene Algebras and the Algebra of ...
    In this paper we introduce yet another: a Kleene algebra is any model of the equations and equational implications given in x2. By general considerations of ...
  66. [66]
    Tropical Semirings
    Apr 28, 2013 · In particular, note that there are no (tropical) additive inverses (except for ), and every element in has a (tropical) multiplicative inverse. ...
  67. [67]
    Minus Infinity | - James Propp
    May 16, 2017 · More formally, algebraists say that the operation of negation gives an isomorphism between the min-plus semiring and max-plus semiring . This is ...Missing: inverse | Show results with:inverse
  68. [68]
    Semirings, Automata, Languages - SpringerLink
    Free delivery 14-day returnsThe purpose of this book is to develop the theory of automata and formal languages, starting from ideas based on linear algebra. ... Werner Kuich, Arto Salomaa.
  69. [69]
  70. [70]
    Introduction to Lattices and Order
    This new edition of Introduction to Lattices and Order presents a radical reorganization and updating, though its primary aim is unchanged.
  71. [71]
    Galois connections and Galois theory
    Jul 24, 2022 · A Galois connection is just a pair of functions between partially ordered sets that have some nice properties.
  72. [72]
    [PDF] The Galois Connection between Syntax and Semantics
    Definition 1.1.1 A partially ordered set P = (P, 4) – henceforth, a poset – is a set P, carrying an ordering relation 4 ...
  73. [73]
    [PDF] Galois Connections for Dependencies in Databases
    Nov 17, 2017 · dependencies in data mining (association rules) are naturally expressed in terms of closed descriptions, Galois connections, and concept (Galois) ...