Fact-checked by Grok 2 weeks ago

Principal ideal domain

In , particularly within , a principal ideal domain (PID) is defined as an —a with unity and no zero divisors—in which every is principal, meaning it can be generated by a single element. This property ensures that the ring's ideal structure is highly organized and simplifies many algebraic constructions./02:_Fields_and_Rings/2.04:_Principal_Ideals_and_Euclidean_Domains) PIDs are significant because they generalize the unique factorization properties of the integers, making them fundamental in and . Every PID is a (UFD), where every non-zero non-unit element factors uniquely into irreducibles up to units and order. Additionally, PIDs are Noetherian rings, satisfying the ascending chain condition on ideals, which aids in studying theory and . Prominent examples of PIDs include the ring of integers , the k over any k, and the ring of Gaussian integers . A key subclass consists of Euclidean domains, such as and k, which admit a and thus are PIDs by virtue of their norm functions enabling the division algorithm./02:_Fields_and_Rings/2.04:_Principal_Ideals_and_Euclidean_Domains) These structures underpin applications in , , and the study of algebraic curves.

Fundamentals

Definition

In , an is defined as a with a (unity) that is not zero and contains no zero divisors, meaning that if the product of two nonzero elements is zero, then at least one of the elements must be zero. A in such a ring R is an generated by a single element a \in R, denoted (a), which consists of all multiples of a by elements of R; formally, (a) = \{ ra \mid r \in R \}. A (PID) is an R in which every I of R is principal, that is, I = (a) for some a \in R. This structure generalizes the \mathbb{Z}, where every ideal takes the form n\mathbb{Z} for some n.

Principal ideals

In an R, a principal ideal generated by an element a \in R is the set (a) = \{ r a \mid r \in R \}, consisting of all multiples of a by elements of the . This construction yields an of R, as it is closed under addition and under multiplication by any element of R. If a = 0, then (0) is the zero ideal \{0\}. A key property of principal ideals is the absorption relation: (a) \subseteq (b) b divides a in R. For operations on principal ideals, the sum (a) + (b) equals (d), where d = \gcd(a, b) provided that a exists in R. The product of principal ideals is (a)(b) = (ab), since the ideal generated by all products of from (a) and (b) is precisely the set of multiples of ab. Principal ideals generated by units or associate elements coincide: if u \in R is a unit, then (u) = R, the entire ring, and more generally, (a) = (b) if and only if a and b are , meaning a = b u for some unit u \in R.

Key Properties

Unique factorization domains

In an integral domain, an element p is called irreducible if it is nonzero and not a unit, and whenever p = ab for elements a, b in the domain, one of a or b must be a unit. An element p is called prime if it is nonzero and not a unit, and whenever p divides the product ab, then p divides a or p divides b. In any , every is irreducible, but the converse holds the domain is a . A (UFD) is an in which every nonzero non-unit element can be written as a product of irreducible elements, and this factorization is unique up to the order of the factors and multiplication by units (i.e., associates). In a UFD, the irreducibles coincide with the primes, ensuring that the prime ideals are precisely those generated by irreducible elements. Every principal ideal domain is a unique factorization domain. To see this, first note that a principal ideal domain is Noetherian, so every nonzero non-unit element admits a factorization into irreducibles: if not, the set of such elements would admit an infinite ascending chain of principal ideals under divisibility, contradicting the Noetherian property. For uniqueness, suppose a = p_1 \cdots p_n = q_1 \cdots q_m are two factorizations into irreducibles; without loss of generality, assume n \leq m. Then (p_1 \cdots p_n) = (q_1 \cdots q_m) as principal ideals, so (q_1 \cdots q_m) is contained in the maximal ideal (p_1), implying p_1 divides some q_i, hence p_1 associates to q_i by irreducibility. Iterating this process shows all factors associate pairwise after reordering. In a principal ideal domain, the unique factorization of elements induces a corresponding of principal s: if a = u p_1^{e_1} \cdots p_k^{e_k} with u a and p_i distinct irreducibles, then the principal (a) = (p_1)^{e_1} \cdots (p_k)^{e_k}, where each (p_i) is prime. This correspondence highlights how the principal ideal structure enforces uniqueness at the ideal level, mirroring element-wise . As motivation for polynomial rings, Gauss's lemma states that if R is a UFD, then the product of two primitive polynomials (those with content 1) in R is primitive, implying that R is also a UFD. This result relies on the unique factorization in R to control the contents of polynomial products.

Euclidean domains and PIDs

A domain is an R equipped with a Euclidean function v: R \setminus \{0\} \to \mathbb{N} \cup \{0\}, where \mathbb{N} denotes the non-negative integers, such that for any a, b \in R with b \neq 0, there exist q, r \in R satisfying a = qb + r and either r = 0 or v(r) < v(b). Every Euclidean domain is a principal ideal domain. To see this, consider a nonzero ideal I in the Euclidean domain R; let d \in I \setminus \{0\} minimize v(d). For any a \in I, apply the division algorithm to obtain a = qd + r with r = 0 or v(r) < v(d); since r = a - qd \in I, the minimality of v(d) forces r = 0, so d generates I. In a Euclidean domain, the greatest common divisor of two elements a, b \neq 0 can be computed via the Euclidean algorithm: repeatedly replace (a, b) with (b, r) where a = qb + r and r = 0 or v(r) < v(b), until the remainder is zero; the last nonzero remainder is a gcd of a and b. Examples of Euclidean functions include the absolute value v(a) = |a| on the ring of integers \mathbb{Z}, and the degree v(f) = \deg(f) on the polynomial ring F over a field F.

Dimensional aspects

The Krull dimension of a commutative ring R is defined as the supremum of the lengths of strictly ascending chains of prime ideals in R, where the length of a chain \mathfrak{p}_0 \subsetneq \mathfrak{p}_1 \subsetneq \cdots \subsetneq \mathfrak{p}_n is n. Every principal ideal domain (PID) has Krull dimension at most 1. If the PID is a field, then its only prime ideal is the zero ideal, yielding dimension 0. For non-field PIDs, such as the ring of integers \mathbb{Z}, the dimension is exactly 1. To see this, note that in a PID, every nonzero prime ideal is principal, generated by a prime element, and such ideals are . The only possible chains of prime ideals are thus of the form (0) \subsetneq (p), where p is a prime element; longer chains are impossible. PIDs satisfy the ascending chain condition (ACC) on principal ideals, as they are , implying stabilization of any ascending chain of principal ideals. Regarding the descending chain condition (DCC), PIDs do not generally satisfy it on principal ideals, as seen in infinite descending chains like (2) \supsetneq (4) \supsetneq (8) \supsetneq \cdots in \mathbb{Z}. However, in the special case where the PID is a field, it satisfies the DCC on all ideals and is thus an .

Examples and Counterexamples

Principal ideal domains

A principal ideal domain (PID) is an integral domain in which every ideal is principal, meaning it can be generated by a single element. The ring of integers \mathbb{Z} serves as the prototypical example of a PID. Every ideal in \mathbb{Z} is of the form (n) for some nonnegative integer n \geq 0, consisting of all integer multiples of n; this includes the zero ideal (0) and the unit ideal (1) = \mathbb{Z}. The principal nature of these ideals follows from the well-ordering principle applied to the nonnegative elements of any nonzero ideal, ensuring a unique generator as the smallest positive element therein. Fields provide another fundamental class of PIDs, as they possess only two ideals: the zero ideal (0) and the entire field k = (1). In a field k, every nonzero element is a unit, so any nonzero ideal must contain 1 and thus coincide with k. This trivial ideal structure underscores why all fields, including finite fields such as \mathbb{Z}/p\mathbb{Z} for prime p, qualify as PIDs. For instance, \mathbb{Z}/p\mathbb{Z} is a finite field with p elements, and its ideals are precisely (0) and itself. Polynomial rings over fields also exemplify PIDs. For any field k, the ring k of polynomials in one indeterminate x has the property that every ideal is principal, generated by a single polynomial. This arises because k admits a Euclidean algorithm based on polynomial division, where the degree function serves as the norm, allowing division with remainder of strictly lower degree. Ideals in k are thus of the form (f(x)) for some f(x) \in k, with monic polynomials often serving as canonical generators. Certain quadratic integer rings further illustrate PIDs among algebraic integers. The Gaussian integers \mathbb{Z} = \{a + bi \mid a, b \in \mathbb{Z}\}, where i = \sqrt{-1}, form a PID equipped with the Euclidean norm N(a + bi) = a^2 + b^2. This norm enables a division algorithm, ensuring every ideal is principal; for example, the ideal generated by 2 factors as (1+i)^2 up to units, but all ideals remain singly generated. Such rings highlight how Euclidean domains, a subclass of PIDs, appear in number theory.

Non-principal ideal domains

A classic example of an integral domain that is not a principal ideal domain (PID) is the polynomial ring k[x, y] in two indeterminates over a field k. This ring is Noetherian but fails to be a PID because the ideal (x, y), generated by x and y, is not principal. To see this, suppose (x, y) = (f) for some f \in k[x, y]. Then f divides both x and y, implying that f is a non-constant polynomial of degree 1 (up to units). However, such an f cannot generate (x, y), because the vector space of degree-1 homogeneous polynomials in (f) is 1-dimensional (spanned by f), whereas in (x, y) it is 2-dimensional (spanned by x and y). In algebraic number theory, rings of integers in quadratic number fields with class number greater than 1 provide further counterexamples. For instance, consider the ring \mathbb{Z}[\sqrt{-5}], the ring of integers of \mathbb{Q}(\sqrt{-5}). This is an integral domain, but the ideal I = (2, 1 + \sqrt{-5}) is not principal. Elements of I are of the form $2a + (1 + \sqrt{-5})b for a, b \in \mathbb{Z}[\sqrt{-5}], and the norm function N(\alpha + \beta \sqrt{-5}) = \alpha^2 + 5\beta^2 shows that no single generator can produce all such elements, as N(I) = 2 while principal ideals have norms that are squares or products accordingly; moreover, \mathbb{Z}[\sqrt{-5}] also fails unique factorization of elements, with $6 = 2 \cdot 3 = (1 + \sqrt{-5})(1 - \sqrt{-5}) providing distinct factorizations into irreducibles. Another example is the ring of integers of \mathbb{Q}(\sqrt{-23}), which has class number 3, implying non-principal ideals exist since the ideal class group is non-trivial. The ring of entire functions, denoted E, consisting of all holomorphic functions on the complex plane \mathbb{C}, is another integral domain that is not a PID. While E is a Bézout domain—meaning every finitely generated ideal is principal—it is not Noetherian, so there exist ideals that are not finitely generated and hence not principal. For example, the ideal of entire functions vanishing on a discrete set with a limit point at infinity cannot be generated by a single element. These examples illustrate common reasons for failure to be a PID: in the polynomial case, the lack of a suitable division algorithm beyond one variable prevents all ideals from being principal; in quadratic integer rings with class number greater than 1, non-trivial ideal classes mean some ideals are not principal; and in the entire functions case, the absence of the ascending chain condition allows infinitely generated ideals. In general, such domains lack a Euclidean function or fail to have unique factorization into principal ideals.

Modules and Applications

Finitely generated modules

Let R be a . An R-module is an abelian group M together with a bilinear map R \times M \to M (scalar multiplication) satisfying the usual axioms: distributivity over addition in R and M, and compatibility with the ring multiplication in R. A module M is finitely generated if there exists a finite set \{ m_1, \dots, m_n \} \subseteq M such that every element of M is a finite R-linear combination of the m_i. Over a PID R, every finitely generated module M admits a unique decomposition M \cong F \oplus T, where F is a free R-module (the free part) and T is the torsion submodule of M (the torsion part). The torsion submodule T = \mathrm{Tor}(M) consists of all torsion elements of M, i.e., elements m \in M such that there exists a nonzero r \in R with r m = 0. It is itself a submodule, and M / T is torsion-free. A finitely generated torsion-free module over a PID is free. A free R-module is isomorphic to a direct sum of copies of R, denoted R^r for some nonnegative integer r, called the rank of the free module. For a torsion-free finitely generated module M, the rank r equals the dimension of the vector space M \otimes_R K over the quotient field K = \mathrm{Frac}(R). The torsion submodule T of a finitely generated module over a PID can be further decomposed in two canonical ways: using invariant factors, where T \cong \bigoplus_{i=1}^m R / (a_i R) with a_1 \mid a_2 \mid \dots \mid a_m and each a_i nonzero and non-unit; or using elementary divisors, where T \cong \bigoplus_j R / (p_j^{e_j} R) for prime elements p_j of R and positive exponents e_j. These decompositions are unique up to ordering and provide complete invariants for the isomorphism class of T.

Structure theorem for modules

The structure theorem for finitely generated modules over a principal ideal domain (PID) provides a complete classification of such modules up to isomorphism. Let R be a PID and M a finitely generated R-module. Then M is isomorphic to a direct sum of the form M \cong R^r \oplus R/(d_1 R) \oplus \cdots \oplus R/(d_k R), where r \geq 0 is the rank of the free part, k \geq 0, and the invariant factors d_1, \dots, d_k are nonzero non-unit elements of R satisfying d_1 \mid d_2 \mid \cdots \mid d_k. This decomposition separates M into its free submodule R^r and its torsion submodule, which is the direct sum of the cyclic torsion modules R/(d_i R). The torsion submodule is zero if and only if M is free. An alternative form of the theorem uses elementary divisors, expressing the torsion part as a direct sum over prime elements p of R: M \cong R^r \oplus \bigoplus_p \bigoplus_{j=1}^{m_p} R/(p^{e_{p,j}} R), where each e_{p,j} > 0 and the exponents are ordered such that e_{p,1} \leq e_{p,2} \leq \cdots \leq e_{p,m_p} for each prime p. The invariant factors and elementary divisors are related: the elementary divisors are obtained by factoring each invariant factor d_i into prime powers, and the invariant factors can be reconstructed by multiplying the highest powers of each prime across the factors. Both forms are unique up to the associates of the generators (i.e., units in R) and the ordering of the summands in the elementary divisor decomposition. The proof proceeds by first decomposing M into a free part and a torsion part, then classifying the finitely generated torsion module via its matrix. Given generators and relations for M, the relation matrix can be transformed using elementary row and column operations over R (permissible since R is a PID) into , a \operatorname{diag}(d_1, \dots, d_k, 0, \dots, 0) where d_i \mid d_{i+1} and the zeros correspond to the free rank r. This diagonal form yields the invariant factor decomposition directly, and the uniqueness follows from the uniqueness of the up to units in R. The elementary divisor form is derived by prime of the diagonal entries.

Invariants and elementary divisors

In the structure theorem for finitely generated modules over a principal ideal domain (PID) R, the torsion submodule admits two canonical decompositions: one into invariant factors and one into elementary divisors. The invariant factors d_1, d_2, \dots, d_k (with d_i non-units in R and d_i \mid d_{i+1}) arise from the Smith normal form of the relation matrix presenting the module. For a relation matrix A \in M_{m \times n}(R), the Smith normal form is obtained via elementary row and column operations (adding multiples of one row/column to another, swapping, and multiplying by units), yielding a diagonal matrix \operatorname{diag}(d_1, d_2, \dots, d_r, 0, \dots, 0) where d_i \mid d_{i+1} and the non-zero d_i are the invariant factors of the cokernel module \coker(A) \cong \bigoplus_{i=1}^r R/(d_i). The invariant factors can be computed without fully reducing to by using determinants of minors: the product d_1 d_2 \cdots d_k equals the (in [R](/page/R)) of all k \times k minors of A, for each k = 1, \dots, \min(m,n); thus, d_1 is the gcd of all $1 \times 1 minors, d_2 is the gcd of all $2 \times 2 minors divided by d_1, and so on (with d_0 = 1). This ensures the divisibility condition and provides the diagonal entries directly. The resulting module is then \bigoplus_{i=1}^k R/(d_i) \oplus [R](/page/R)^{n - r}, where r is the of the torsion-free part. Elementary divisors provide an alternative primary decomposition of the torsion submodule. For each prime element p \in R, the p-primary component decomposes as \bigoplus_j R/(p^{e_j}) where $0 < e_1 \leq e_2 \leq \cdots \leq e_m are the exponents (unique up to ordering), and the full torsion module is the direct sum over all such p of these components. To obtain the elementary divisors from the invariant factors, factor each d_i = u_i \prod_p p^{a_{i,p}} (with unit u_i) into its prime power factors, then for each p, collect and sort the exponents \{a_{i,p} \mid i=1,\dots,k, a_{i,p} > 0\} in non-decreasing order to get the e_j for that p. Conversely, to recover invariant factors from elementary divisors, align the exponent lists for each p by decreasing order (padding shorter lists with zeros), then form d_k as the product over p of p raised to the k-th exponent in that list (from the end), ensuring d_1 \mid d_2 \mid \cdots \mid d_l where l is the maximum length over all p-lists. This correspondence relies on the Chinese Remainder Theorem for coprime ideals. Consider the \mathbb{[Z](/page/Z)}-module presented by generators x, y and relations $2x = 0, $4y = 0, given by the relation matrix A = \begin{pmatrix} 2 & 0 \\ 0 & 4 \end{pmatrix}. The $1 \times 1 minors are $2, 0, 0, 4, with \gcd(2,4) = 2 = d_1. The sole $2 \times 2 minor is \det(A) = 8, so d_2 = 8 / 2 = 4. Thus, the invariant factors are $2, 4, and the module is \mathbb{Z}/2\mathbb{Z} \oplus \mathbb{Z}/4\mathbb{Z}. Factoring gives elementary divisors $2^1 (from $2) and $2^2 (from $4), so the module is also \mathbb{Z}/2\mathbb{Z} \oplus \mathbb{Z}/4\mathbb{Z} in elementary form. Two finitely generated torsion modules over a PID are isomorphic if and only if their factors coincide (up to units) or, equivalently, their elementary divisors coincide (as multisets). This uniqueness follows from the uniqueness of the and the . A key application is the of finite abelian groups as torsion \mathbb{Z}-modules over the PID \mathbb{Z}, where the factors yield a decomposition into cyclic groups of orders d_1 \mid d_2 \mid \cdots \mid d_k and the elementary divisors into cyclic p-groups, facilitating computations like the order or structure of group extensions.

Advanced Relations

Dedekind domains

A is defined as an integrally closed Noetherian of 1, meaning that every nonzero is maximal. This structure generalizes (PIDs) by relaxing the requirement that all be principal while preserving a form of unique factorization at the level of ideals. In a Dedekind domain, every nonzero proper ideal factors uniquely into a product of prime ideals, up to ordering and units; this characterization holds precisely for integrally closed Noetherian domains satisfying this unique factorization property for ideals. Principal ideal domains form a special subclass of Dedekind domains where every ideal is principal, which is equivalent to the ideal class group being trivial. The ideal class group of a Dedekind domain R with field of fractions K is the quotient group of the multiplicative group of fractional ideals of R by the subgroup of principal fractional ideals; its order is called the class number h(R). Thus, a Dedekind domain is a PID if and only if its class number is 1, ensuring that unique factorization extends from ideals to elements. Rings of integers in number fields provide canonical examples of Dedekind domains. For instance, the ring of integers \mathcal{O}_K of a number field K is always a Dedekind domain, and it is a PID precisely when the class number h(K) = 1. The Gaussian integers \mathbb{Z}, which are the ring of integers of \mathbb{Q}(\sqrt{-1}), form such a PID since \mathbb{Q}(\sqrt{-1}) has class number 1.

Connections to algebraic geometry

In algebraic geometry, principal ideal domains (PIDs) emerge as coordinate rings of specific affine varieties over a k. For an V \subseteq \mathbb{A}^n_k, the coordinate ring is k[V] = k[x_1, \dots, x_n]/I(V), where I(V) denotes the vanishing ideal of V. This ring is a PID when V is the \mathbb{A}^1_k, as I(\mathbb{A}^1_k) = (0) and k[\mathbb{A}^1_k] \cong k, the in one , which is a well-known PID. Similarly, for a point V = \{a\} \subseteq \mathbb{A}^n_k with a = (a_1, \dots, a_n), the ideal I(V) = (x_1 - a_1, \dots, x_n - a_n) is maximal, yielding k[V] \cong k, a and hence a PID. In these cases, the geometry simplifies to zero- or one-dimensional objects where all ideals are principal. The polynomial ring k as a PID directly corresponds to the affine line, a curve of genus zero whose rational structure allows every in its coordinate ring to be generated by a single element. More generally, the coordinate ring of any affine of zero is a PID, reflecting the tame divisor theory and unique factorization on such varieties. Hilbert's Nullstellensatz provides key implications for PIDs in this context, particularly over algebraically closed fields. It establishes that maximal ideals in k[x_1, \dots, x_n] are precisely those of the form (x_1 - a_1, \dots, x_n - a_n) for points a \in k^n, and in the one-variable case k, these are principal ideals (x - a). This correspondence underscores how PIDs like k capture the maximal ideals geometrically as points on the affine line, linking algebraic structure to points via the Nullstellensatz. PIDs also connect to normalization in geometric settings, as they are integrally closed domains—hence rings—meaning their fraction fields contain no elements over the ring outside the ring itself. In algebraic geometry, the of an affine variety produces an integrally closed coordinate ring, and PIDs exemplify such domains for simple varieties like the affine line, where the ring is already normal without needing normalization. In the projective setting, PIDs appear less frequently due to the use of homogeneous ideals in the graded coordinate ring k[x_0, \dots, x_n] of projective space \mathbb{P}^n_k. While hypersurfaces in \mathbb{P}^n_k are defined by principal homogeneous ideals (generated by a single irreducible homogeneous polynomial), the resulting quotient rings are rarely PIDs beyond low dimensions, as higher-dimensional projective varieties introduce non-principal ideals in their affine cones. Thus, PIDs remain tied to affine rather than projective geometries in most cases.

Historical development

The concept of unique factorization in the integers, as established by in his (1801), laid the groundwork for exploring similar properties in more general number rings, including quadratic integer rings like the Gaussian integers \mathbb{Z}. Gauss demonstrated that these rings often admit unique factorization up to units, but this property fails in certain quadratic fields, such as \mathbb{Z}[\sqrt{-5}], where $6 = 2 \cdot 3 = (1 + \sqrt{-5})(1 - \sqrt{-5}) shows non-unique factorizations into irreducibles. This failure highlighted the limitations of classical and motivated the search for structures preserving factorization properties, leading toward the development of ideal theory in the mid-19th century. Ernst Kummer, in the , introduced "ideal numbers" to address issues in cyclotomic fields, providing a precursor to modern ideals by treating complex factors as abstract entities to restore ness. Building on this, refined the approach in his supplements to Dirichlet's Vorlesungen über Zahlentheorie (first edition 1871), defining ideals as sets closed under addition and multiplication by ring elements, and proving that in rings of algebraic integers, every nonzero ideal factors ly into prime ideals. Dedekind recognized principal ideal domains (PIDs) as the special case where all ideals are principal, generalizing the of integers and linking it to Euclidean-like algorithms in domains such as \mathbb{Z} and rings. These developments, spanning the to 1870s, shifted focus from elements to ideals, establishing PIDs as a cornerstone of . In the 1890s, advanced the theory in his Zahlbericht (1897), where he explored over rings of integers in number fields—often PIDs—and proved results on the finite generation and structure of ideal classes, including theorems on module bases in fields that underscored the role of PIDs in classifying abelian extensions. This work bridged concrete number-theoretic to more abstract linear algebra, emphasizing finite bases for torsion-free over PIDs. The early 20th century saw abstract these ideas beyond number fields, introducing general commutative rings and Noetherian conditions in her 1921 paper on ideals and her 1927 work on hypercomplex systems, which formalized PIDs as integral domains where every ideal is singly generated. Concurrently, the notion of Euclidean domains—rings admitting a , implying they are PIDs—was explicitly termed and characterized , drawing from earlier examples like \mathbb{Z} and \mathbb{Z}. , in his 1920s lectures and papers on noncommutative algebras, further developed over such domains, proving theorems for representations. By the mid-20th century, texts like Atiyah and Macdonald's Introduction to (1969) solidified PIDs' central role in modern , integrating them into broader .