Fact-checked by Grok 2 weeks ago

Zero divisor

In , particularly within the theory of , a zero divisor is a nonzero a in a ring R such that there exists another nonzero b \in R with a \cdot b = [0](/page/0) or b \cdot a = [0](/page/0), where \cdot denotes the ring's . More precisely, if the multiplication is non-commutative, one distinguishes between left zero divisors (where a \cdot b = [0](/page/0) for some nonzero b) and right zero divisors (where b \cdot a = [0](/page/0) for some nonzero b); in commutative rings, these coincide. The zero element itself is trivially a zero divisor in any nontrivial ring, but the concept focuses on nonzero instances, which prevent the ring from satisfying the zero product property—wherein the product of two nonzero elements is nonzero. Commutative rings with multiplicative identity lacking nonzero zero divisors are termed integral domains, a fundamental class that includes the integers \mathbb{Z} and polynomial rings over fields, enabling key theorems like unique factorization. Classic examples abound in quotient rings: in \mathbb{Z}/6\mathbb{Z}, the elements 2 and 3 are zero divisors since $2 \cdot 3 = 0 \pmod{6}, yet neither is zero. Similarly, in the ring of $2 \times 2 matrices over the reals, the matrices \begin{pmatrix} 1 & 0 \\ 0 & 0 \end{pmatrix} and \begin{pmatrix} 0 & 0 \\ 0 & 1 \end{pmatrix} multiply to the zero matrix. Zero divisors play a crucial role in , influencing ideal structure, localization, and the study of modules, as their presence signals deviations from domain-like behavior essential for and .

Definition and Fundamentals

Definition

In , a is an consisting of a set R equipped with two binary operations, and , such that (R, +) forms an , is associative, and the distributive laws hold: a(b + c) = ab + ac and (a + b)c = ac + bc for all a, b, c \in R. A nonzero a \in R is called a left zero divisor if there exists a nonzero b \in R such that a b = [0](/page/0), where [0](/page/0) denotes the of the . Similarly, a nonzero element a \in R is a right zero divisor if there exists a nonzero element b \in R such that b a = 0. An element that is both a left zero divisor and a right zero divisor is called a two-sided zero divisor. In a , where multiplication satisfies a b = b a for all a, b \in R, the notions of left, right, and two-sided zero divisors coincide./02%3A_Fields_and_Rings/2.02%3A_Rings) Rings without nonzero zero divisors are known as integral domains.

Relation to Units and Integral Domains

In ring theory, units are elements that possess multiplicative inverses, meaning for an element u in a ring R, there exists u^{-1} \in R such that u u^{-1} = u^{-1} u = 1, where $1 is the multiplicative identity. A zero divisor cannot be a unit, as the existence of a nonzero element b such that a b = 0 leads to a contradiction if a is assumed invertible: multiplying both sides by a^{-1} yields b = a^{-1} (a b) = a^{-1} 0 = 0. This distinction underscores how zero divisors disrupt the invertibility essential to units, preventing rings with zero divisors from exhibiting field-like behavior for all nonzero elements. Integral domains represent a class of rings free from such disruptions. Specifically, an is defined as a with (where the unity $1 \neq 0) that contains no zero divisors other than zero itself, ensuring that the product of any two nonzero elements is nonzero./16:_Rings/16.04:_Integral_Domains_and_Fields) Equivalently, a with is an if and only if it has no zero divisors, as this condition directly enforces the absence of nontrivial zero products. This characterization classifies rings based on the presence or absence of zero divisors, with integral domains serving as a foundational structure analogous to the integers \mathbb{Z}, where divisibility behaves predictably without zero-induced collapses. Fields extend this structure further, forming integral domains in which every nonzero element is a . In a , the lack of zero divisors combined with universal invertibility for nonzero elements ensures full division capability, distinguishing fields as the maximal integral domains under this criterion./16:_Rings/16.04:_Integral_Domains_and_Fields) Thus, while all fields are integral domains, the converse holds only when invertibility permeates the entire nonzero , highlighting the hierarchical relationship between zero divisors, units, and domain classifications.

Examples and Non-Examples

Concrete Examples

A concrete illustration of zero divisors appears in the ring of 2×2 matrices over \mathbb{Z}/6\mathbb{Z}, denoted M_2(\mathbb{Z}/6\mathbb{Z}). Consider the matrices A = \begin{pmatrix} 2 & 0 \\ 0 & 3 \end{pmatrix} and B = \begin{pmatrix} 3 & 0 \\ 0 & 2 \end{pmatrix}. Their product is AB = \begin{pmatrix} 6 & 0 \\ 0 & 6 \end{pmatrix} = \begin{pmatrix} 0 & 0 \\ 0 & 0 \end{pmatrix}, the zero matrix in this ring, while both A and B are nonzero. In the polynomial ring (\mathbb{Z}/6\mathbb{Z}), constant polynomials and nonconstant ones can serve as zero divisors when the coefficient ring has them. For instance, the constant polynomial $3 (viewed as $3 + 0x + \cdots) and the linear polynomial $2x satisfy $3 \cdot 2x = 6x = 0 \cdot x = 0, with both nonzero in the ring. This follows from the general criterion that a polynomial is a zero divisor if the greatest common divisor of its coefficients and $6 exceeds $1. Direct products of rings with zero divisors also exhibit them prominently. In \mathbb{Z}/6\mathbb{Z} \times \mathbb{Z}/6\mathbb{Z} under componentwise and , the elements (2, [0](/page/0)) and (3, [0](/page/0)) are nonzero, but their product is (2 \cdot 3, [0](/page/0) \cdot [0](/page/0)) = (6, [0](/page/0)) = ([0](/page/0), [0](/page/0)), the . This pattern generalizes: in a R \times S, elements like (r, [0](/page/0)) and (0, s) with r, s nonzero yield zero when multiplied, provided R and S are nonzero rings. Rings of functions from a set to a ring, with pointwise operations, provide another intuitive class of examples. Consider the ring of all functions f: X \to \mathbb{Z}/6\mathbb{Z} where X has at least two elements, say X = \{1, 2\}, under pointwise addition and multiplication. Let \chi_1 be the characteristic function with \chi_1(1) = 1 and \chi_1(2) = 0, and \chi_2 with \chi_2(1) = 0 and \chi_2(2) = 1. Then \chi_1 \cdot \chi_2 is the zero function, as it evaluates to $0 at both points, while both \chi_1 and \chi_2 are nonzero. More generally, functions with disjoint supports multiply to the zero function in such rings.

Non-Examples in Common Rings

The ring of integers \mathbb{Z} provides a fundamental non-example of a ring containing zero divisors. In \mathbb{Z}, the product of two nonzero elements is always nonzero, as \mathbb{Z} is an where ab = 0 implies a = 0 or b = 0. This property follows directly from , which states that if a prime p divides the product ab, then p divides a or p divides b, ensuring no nontrivial zero divisors exist. Polynomial rings over fields similarly lack zero divisors. For a field k, the ring k of polynomials in one indeterminate x is an integral domain, meaning the product of two nonzero polynomials is nonzero. This absence of zero divisors is guaranteed by Gauss's lemma, which asserts that the content of a product of primitive polynomials is the product of their contents, leading to unique factorization and precluding zero divisors in such rings. Fields themselves, such as the rational numbers \mathbb{Q} or the real numbers \mathbb{R}, contain no zero divisors by definition. In a field, every nonzero element has a , so if ab = 0 with a \neq 0, then multiplying both sides by a^{-1} yields b = 0. These structures—\mathbb{Z}, k for fields k, and fields—are all integral domains, serving as essential building blocks in for constructing quotient rings, modules, and other extensions without the disruptions caused by zero divisors.

Algebraic Properties

Basic Properties

In a ring R, let Z(R) denote the set of zero divisors. The set Z(R) is not necessarily closed under addition, as there exist rings where the sum of two zero divisors is neither zero nor a zero divisor; for example, in the ring \mathbb{R} \times \mathbb{R}, the elements (1,0) and (0,1) are zero divisors, but their sum (1,1) is a . However, Z(R) exhibits closure properties under : the product of a non-zero element and a zero divisor is either zero or itself a zero divisor. Specifically, if a \in Z(R) with a \neq 0 and there exists b \neq 0 such that ab = 0, then for any r \in R with r \neq 0, (ra)b = r(ab) = 0, so ra annihilates b \neq 0; if additionally ra \neq 0, then ra \in Z(R). In a R, if a \in Z(R) is a zero divisor, then any non-zero multiple ka for k \in R \setminus \{0\} with ka \neq 0 is also a zero divisor. To see this, suppose ab = 0 for some b \neq 0; then (ka)b = k(ab) = 0, so ka annihilates the non-zero b. This property highlights the "propagation" of zero-divisor behavior under scaling by non-zero elements in commutative settings. The presence of zero divisors in a prevents the validity of cancellation laws for . In particular, if R has a zero divisor a \neq 0, there exist b, c \in R with b \neq c but ab = ac; for instance, if ab = 0 = ac with b \neq c and both non-zero. Conversely, a ring satisfies the cancellation laws—if ab = ac and a \neq 0 implies b = c— it has no zero divisors. This equivalence underscores the role of zero divisors in obstructing unique factorization and division-like operations. For a non-zero element a \in R, the (left) annihilator is defined as \operatorname{Ann}(a) = \{x \in R \mid ax = 0\}. If a is a zero divisor, then \operatorname{Ann}(a) \neq \{0\}, since there exists some non-zero b with ab = 0, so b \in \operatorname{Ann}(a). In commutative rings, the annihilator is two-sided, and zero divisors are precisely the non-zero elements with non-trivial annihilators. This concept links zero divisors to theory, as \operatorname{Ann}(a) is always an . Regarding ideals, the set of left zero divisors in a general does not always form a left , as it may fail closure under : if a and a' are left zero divisors with annihilators generated by distinct non-zero elements whose product vanishes, the sum a + a' may not annihilate any non-zero element. However, in commutative Noetherian , the set of zero divisors coincides with the union of the .

Properties in Specific Ring Structures

In quotient rings, the presence of zero divisors is closely tied to the nature of the ideal. Consider a commutative ring R with an ideal I. The quotient ring R/I consists of cosets \bar{r} = r + I for r \in R, with multiplication defined by \bar{a} \bar{b} = \overline{ab}. Thus, \bar{a} \bar{b} = \bar{0} if and only if ab \in I. Zero divisors in R/I arise from non-zero cosets \bar{a}, \bar{b} (i.e., a \notin I, b \notin I) such that ab \in I; these may lift from zero divisors in R (where ab = 0 \subseteq I) or from elements whose products lie in I without either being in I, such as when one annihilates a non-zero element of I. A key theorem states that R/I is an integral domain (hence has no zero divisors) if and only if I is a prime ideal; therefore, if I is not prime, R/I contains zero divisors, assuming R/I \neq 0. For direct products of rings, zero divisors emerge prominently even when the component rings lack them. Given commutative rings R and S with unity, the direct product R \times S has componentwise addition and multiplication, so elements are pairs (r, s) with r \in R, s \in S. If both R and S are non-trivial (i.e., not the zero ring), then (1_R, 0_S) and (0_R, 1_S) are non-zero elements satisfying (1_R, 0_S) \cdot (0_R, 1_S) = (0_R, 0_S), making them zero divisors. More generally, R \times S always contains zero divisors unless at least one of R or S is the zero ring, as the construction inherently produces such pairs regardless of whether R or S individually has zero divisors. In local rings, the structure of the maximal ideal highlights potential zero divisors among non-units. A commutative ring A is local if it has a unique maximal ideal m, which comprises all non-units of A (i.e., elements outside m are units). While local rings may be integral domains (with no zero divisors), in general, the maximal ideal m can contain zero divisors; for instance, if A is not a domain, elements in m often annihilate non-zero elements within A, placing them among the non-units. This containment underscores how zero divisors, when present, reside in the non-unit ideal m, distinguishing local rings from fields (where m = \{0\}).

Special Cases

The Zero Element as a Zero Divisor

In any R, the $0 satisfies the equation $0 \cdot b = 0 for every element b \in R, including all non-zero b. Thus, by a literal interpretation of the condition for a zero divisor—without the non-zero stipulation—$0 qualifies as both a left and right zero divisor, since there exist non-zero elements b such that $0 \cdot b = 0 = b \cdot 0./02:_Fields_and_Rings/2.02:_Rings) Standard definitions in , however, explicitly exclude the zero by requiring a zero divisor to be a non-zero a \in R such that there exists a non-zero b \in R with a \cdot b = [0](/page/0) or b \cdot a = [0](/page/0). This convention originated in early abstract treatments of , such as Abraham Fraenkel's work on zero divisors and ring decompositions, and was reinforced in Emmy Noether's 1921 paper on theory in ring domains, where the focus on non-zero elements facilitated the study of and structures in non-commutative settings. Including the zero element as a zero divisor would trivialize the concept, as every (except the ) possesses $0 with this property, eliminating its utility in classifying rings like integral domains, which are defined precisely as commutative rings with unity having no zero divisors. The remains the unique absorbing element under in any , but it is not considered a proper zero divisor under this exclusionary convention.

One-Sided Zero Divisors

In non-commutative s, zero divisors exhibit , distinguishing left and right variants from the two-sided case prevalent in commutative settings. A non-zero element a \in [R](/page/R) of a R is a left zero divisor if there exists a non-zero b \in [R](/page/R) such that ab = 0; it is a right zero divisor if there exists a non-zero b \in [R](/page/R) such that ba = 0. An element satisfying both conditions is a two-sided zero divisor. These definitions extend the standard notion of zero divisors, capturing the directional nature of in non-commutative structures. In commutative rings, left and right zero divisors coincide due to the ab = ba, making all zero divisors two-sided by default. However, non-commutativity allows rings where an element is a left zero divisor without being a right zero divisor (or ), illustrating how the lack of commutativity introduces one-sided behavior. This distinction is absent in commutative rings but fundamental to understanding more general algebraic structures, such as rings over fields, where zero divisors often arise from deficiencies in products. A concrete example occurs in the \mathbb{Z}_2- R generated by indeterminates x_i (i \in \mathbb{N}) subject to the relations x_i x_j = 0 for all i < j. Here, x_1 serves as a left zero divisor since x_1 x_2 = 0 with x_2 \neq [0](/page/0). Yet, x_1 is not a right zero divisor, as the right of x_1 contains only the —no non-zero b \in R satisfies b x_1 = 0. This demonstrates a with asymmetric zero divisors, where left-sided annihilation is possible without a corresponding right-sided counterpart. In rings with unity, the existence of a one-sided zero divisor does not generally imply it is two-sided. For instance, the example above is a with unity (adjoining 1 if necessary preserves the relations), yet x_1 remains strictly left-sided. However, in specific classes of s with unity, such as eversible rings—where every left zero divisor is also a right zero divisor and conversely—one-sided zero divisors effectively become two-sided. This property holds in reversible rings (a subclass of eversible rings) but fails in general non-commutative settings, emphasizing the need for additional structural assumptions.

Generalizations and Extensions

Zero Divisors in Modules

In module theory, the concept of zero divisors extends from to the action of a on a . Let R be a and M an R-. An element r \in R is called a zero divisor on M if there exists a nonzero element m \in M such that r m = 0. This generalizes the notion of zero divisors in , where the special case M = R recovers the standard ring-theoretic definition. The set of all elements in R that annihilate the entire module M forms the annihilator ideal \operatorname{Ann}_R(M) = \{ r \in R \mid r M = 0 \}, which is an ideal of R consisting precisely of those zero divisors on M that act trivially on every element of M. A module M is faithful if \operatorname{Ann}_R(M) = 0, meaning no nonzero element of R acts as a zero divisor on the whole M, so the natural map R \to \operatorname{End}_R(M) is injective. In contrast, torsion modules highlight the pervasive role of zero divisors: an R-module M is torsion if every nonzero m \in M has a nonzero annihilator \operatorname{Ann}_R(m) = \{ r \in R \mid r m = 0 \} \neq 0, i.e., some nonzero r \in R acts as a zero divisor on m. For submodules, the zero divisors behave inclusion-wise. If N is a submodule of M, then the set of zero divisors on N is a of those on M: any r \in R that annihilates a nonzero of N also annihilates a nonzero of M. More precisely, the satisfies \operatorname{Ann}_R(N) \subseteq \operatorname{Ann}_R(M), since N \subseteq M implies that elements killing N kill a of M. This containment reflects the hierarchical structure of actions under zero divisors.

Zero Divisors in Non-Commutative Settings

In non-commutative , the concept of zero divisors extends beyond commutative settings, where left and right zero divisors may differ, but the focus here is on their implications for and advanced algebraic properties. , or skew fields, are fundamental examples of non-commutative with no nonzero zero divisors: every nonzero element admits a left and right , generalizing the of while allowing non-commutative multiplication. This absence of zero divisors ensures that the ring is a , and every finite division ring is in fact commutative, hence a , by Wedderburn's little theorem. A key characterization in Artinian rings states that any Artinian ring with no zero divisors is a division ring, reflecting the rigid structure imposed by the descending chain condition on ideals. This result follows from the fact that such rings are semisimple Artinian, decomposing into matrix rings over division rings, but the lack of zero divisors forces the decomposition to a single division ring component. More generally, non-commutative domains (rings without zero divisors) satisfying the Ore condition embed into their classical ring of right quotients, which is a division ring, extending the framework to non-Artinian cases. In non-commutative , orders—full subrings of finite in the —often inherit or exhibit zero divisor properties from the ambient structure. For instance, the non-commutative (or ) over an , such as the free k\langle x_1, \dots, x_n \rangle in non-commuting indeterminates over a k, is itself a with no zero divisors, as products of nonzero are nonzero due to the free generation and the base 's . However, in more general non-commutative over rings with zero divisors, such as rings, zero divisors propagate, leading to complex behaviors distinct from commutative cases. Unlike commutative rings, where the presence of zero divisors implies the ring is not prime (as the zero ideal would not be prime), non-commutative zero divisors can occur in prime rings without violating primeness. A prime non-commutative ring has no two nonzero two-sided ideals I, J such that IJ = 0, yet it may contain zero divisors if they do not generate such annihilating ideal pairs; examples include certain simple Artinian rings or enveloping algebras with specific zero divisor configurations that preserve ideal products. This distinction highlights how non-commutativity allows for richer ideal structures where zero divisors do not immediately force non-primeness. Modern applications leverage zero divisors to detect structural degeneracy in non-commutative settings. In algebras, absolute zero divisors—nonzero elements x such that [x, [x, L]] = 0—signal degeneracy, as non-degenerate Lie algebras have no such nonzero elements, ensuring the is faithful in the sense of no nonzero elements of index 2; simple Lie algebras over algebraically closed fields of characteristic zero, for example, lack such divisors unless exceptional. Similarly, in operator algebras, positive zero divisors in C*-algebras serve as tools to construct hereditary C*-subalgebras and invariants, detecting non-trivial ideals or non-simplicity that indicate degeneracy in the or K-theoretic properties of the algebra.

References

  1. [1]
    Zero Divisor -- from Wolfram MathWorld
    A nonzero element x of a ring for which x·y=0 , where y is some other nonzero element and the multiplication x·y is the multiplication of the ring.
  2. [2]
    zero-divisor in nLab
    Jan 18, 2025 · A zero-divisor is something that, like zero itself, when multiplied by something possibly nonzero still produces zero as a product. 2.Definitions · In constructive mathematics · Properties
  3. [3]
    Integral Domain -- from Wolfram MathWorld
    Integral Domain. A ring that is commutative under multiplication, has a ... Created, developed and nurtured by Eric Weisstein at Wolfram Research.Missing: coined | Show results with:coined
  4. [4]
    Ring -- from Wolfram MathWorld
    A ring in the mathematical sense is a set S together with two binary operators + and * (commonly interpreted as addition and multiplication, respectively)
  5. [5]
    Zero Divisor - an overview | ScienceDirect Topics
    In general (for a commutative or non-commutative ring), an element that is both a left zero divisor and a right zero divisor is simply called a zero divisor.
  6. [6]
    [PDF] IMMERSE 2008: Assignment 1 - OSU Math
    1.2) Problems involving zerodivisors: (a) Prove that a unit element of a ring cannot be a zerodivisor. Solution. Let a be a unit element and a zero divisor.
  7. [7]
    [PDF] DUMMIT AND FOOTE NOTES Contents 1. Chapter 1: Intro to groups ...
    The set of units in R is R∗. In fact, R∗ is a group. For a field F, we have F∗ = F \ {0}. Note that a zero divisor can never be a unit: let a ∈ R be a ...
  8. [8]
    [PDF] Section IV.19. Integral Domains
    Jul 14, 2023 · Definition 19.6. An integral domain D is a commutative ring with unity 1 6= 0 and containing no divisors of 0. Note 19.A. An integral domain ...
  9. [9]
    [PDF] Math 403 Chapter 13: Integral Domains and Fields 1. Introduction
    An integral domain is a commutative ring with unity that has no zero-divisors. Fields are a subset of integral domains, and every field is an integral domain.
  10. [10]
    [PDF] Contents 2 Rings - Evan Dummit
    A unit can never be a zero divisor in R. ◦ Proof: Suppose first that a is a unit and that xa = 0 for some x 6= 0. ◦ Then by assumption, there is a b such ...
  11. [11]
    [PDF] Nilpotents, units, and zero divisors for polynomials - Keith Conrad
    Example 2.8. A zero divisor in (Z/6Z)[x] must have all coefficients divisible by 2 or all coefficients divisible by 3. Theorem 3.1. A polynomial in A[x1,...,xd] ...
  12. [12]
  13. [13]
    [PDF] Divisibility and Principal Ideal Domains Divisibility. Suppose that R ...
    Suppose that R is an integral domain with unit. If a, b ∈ R, then we ... (Euclid's Lemma.) Suppose that p is an irreducible element of R and that a ...
  14. [14]
    [PDF] The Gauss Lemma and The Eisenstein Criterion
    It is often useful to combine the Gauss Lemma with Eisenstein's criterion. Theorem 2 (Eisenstein) Suppose A is an integral domain and Q ⊂ A is a prime ideal.
  15. [15]
    [PDF] Abstract Algebra II - Auburn University
    Apr 25, 2019 · Q, R, C are integral domains. In fact, every field is an integral ... We need a generalization of these building blocks to our arbitrary integral ...
  16. [16]
    A note on zero-divisors of commutative rings - SpringerLink
    Apr 11, 2012 · Note that if the set of zero-divisors of a ring R forms an ideal I, then I is a ring without identity consisting of zero-divisors. Thus it ...
  17. [17]
    [PDF] M7210 Lecture 28 Friday October 26, 2012 Commutative Rings III
    Oct 26, 2012 · Fact. Every field is an integral domain. Proof . All non-zero elements of a field are units, so there are no zero-divisors.<|separator|>
  18. [18]
    [PDF] Chapter 6, Ideals and quotient rings
    We want a condition on the ideal to make sure there are no zero divisors in the quotient ring. We have seen several examples in the case of principal ideals ...
  19. [19]
    [PDF] 3. Prime and maximal ideals 3.1. Definitions and Examples ...
    A ring A is local if and only if the set of non-units in A form an ideal in A. Proof. Suppose that A is local. Then since every non-unit lies in a maximal ideal.
  20. [20]
    [PDF] THE ORIGINS OF THE DEFINITION OF ABSTRACT RINGS
    Fraenkel's 1914 paper deals mainly with the factorization properties of divisors of zero and with additive decompositions of the elements of the abstract rings ...
  21. [21]
    [PDF] Math 6310. Algebra
    May 20, 2020 · An element a ∈ R is a zero divisor if ab = 0 for some nonzero b ∈ R. As an example, [2] ∈ Z/6Z is a zero divisor. Definition 2.14.6. We ...
  22. [22]
    [PDF] Basic Properties of Rings
    Note that in a division ring every non-zero element a is a unit (since if R ... A nonzero element a in a commutative ring R is called a zero divisor if there ...<|separator|>
  23. [23]
    [PDF] Foundations of Module and Ring Theory
    The correlation of (internal) properties of the module M with properties of the category σ[M] enables a homological classification of modules. Among other ...<|control11|><|separator|>
  24. [24]
    (PDF) Faithful Torsion Modules and Rings - ResearchGate
    Jan 8, 2015 · A module is called faithful if 0 is the only ring element which annihilates all. members of the module. Considered independently, both torsion ...
  25. [25]
    [PDF] Four Group-theoretic Proofs of Wedderburn's Little Theorem - OU Math
    Wedderburn proved in 1905 that a finite division ring is always a field. His result has intrigued generations of mathematicians, spurring generalizations and.<|control11|><|separator|>
  26. [26]
    Background Material on Rings and Modules
    (1) If k is equal to either Q or R, then the ring of quaternions is a division ring. ... (4) If R is an artinian domain, then R is a division ring. Exercise 2.5.
  27. [27]
    [PDF] arXiv:2211.06640v1 [math.RA] 12 Nov 2022
    Nov 12, 2022 · By [20, Corollary], any simple Lie algebra is either a form of a classical Lie algebra, or possesses absolute zero divisors, i.e., nonzero ...
  28. [28]
    [1301.3129] A Note on Positive Zero Divisors in C* Algebras - arXiv
    Jan 13, 2013 · Abstract:In this paper we concern with positive zero divisors in C^{*} algebras. By means of zero divisors, we introduce a hereditary ...