Fact-checked by Grok 2 weeks ago

Quadratic integer

In number theory, quadratic integers are algebraic integers of degree two, specifically the integral elements within quadratic fields, which are finite extensions of the rational numbers \mathbb{Q} obtained by adjoining the square root of a square-free integer D \neq 0,1. These elements take the form \alpha = x + y \sqrt{D} where x, y \in \mathbb{Q}, satisfying a monic quadratic polynomial with integer coefficients, and form the ring of integers \mathcal{O}_K in the quadratic field K = \mathbb{Q}(\sqrt{D}). Quadratic fields are classified as real (when D > 0) or imaginary (when D < 0), with the discriminant D_K determining the structure of \mathcal{O}_K. The ring of integers \mathcal{O}_K is a free \mathbb{Z}-module of rank 2, generated as \mathbb{Z} + \mathbb{Z} \delta where \delta depends on D \mod 4: if D \equiv 2 or $3 \pmod{4}, then \delta = \sqrt{D} and D_K = 4D, so \mathcal{O}_K = \mathbb{Z}[\sqrt{D}]; if D \equiv 1 \pmod{4}, then \delta = \frac{1 + \sqrt{D}}{2} and D_K = D, so \mathcal{O}_K = \mathbb{Z}\left[\frac{1 + \sqrt{D}}{2}\right]. This structure ensures \mathcal{O}_K is the maximal order in K, containing all algebraic integers of the field. Notable examples include the Gaussian integers \mathbb{Z} for D = -1 and the Eisenstein integers \mathbb{Z}[\omega] for D = -3, where \omega = \frac{-1 + \sqrt{-3}}{2}. Key properties of quadratic integers include the norm N(\alpha) = \alpha \overline{\alpha} for \alpha = a + b\sqrt{D}, given by N(\alpha) = a^2 - D b^2, which is a multiplicative homomorphism from \mathcal{O}_K to \mathbb{Z} and measures size in the ring. The conjugate \overline{\alpha} = a - b\sqrt{D} facilitates this, and units are elements \epsilon \in \mathcal{O}_K with N(\epsilon) = \pm 1: for imaginary quadratic fields (D < 0), these are finite groups (e.g., \{\pm 1, \pm i\} for \mathbb{Z}); for real quadratic fields (D > 0), they form infinite cyclic groups generated by a fundamental unit \varepsilon_K. While unique factorization of elements often fails in \mathcal{O}_K (e.g., in \mathbb{Z}[\sqrt{-5}], where $6 = 2 \cdot 3 = (1 + \sqrt{-5})(1 - \sqrt{-5})), ideals factor uniquely into prime ideals, classified as inert, ramified, or based on the modulo primes. These properties underpin applications in , including class numbers and quadratic forms.

Fundamentals

Definition

A quadratic field is a field extension K = \mathbb{Q}(\sqrt{d}) of the rational numbers \mathbb{Q}, where d is a square-free integer not equal to 0 or 1. Algebraic integers are complex numbers that are roots of monic polynomials with integer coefficients. Quadratic integers are the algebraic integers contained in a quadratic field K = \mathbb{Q}(\sqrt{d}). The quadratic integers in K = \mathbb{Q}(\sqrt{d}) form the ring of integers \mathcal{O}_K, which is the integral closure of \mathbb{Z} in K. As a free \mathbb{Z}-module of rank 2, \mathcal{O}_K has an integral basis that depends on the congruence class of d modulo 4: if d \equiv 2 or $3 \pmod{4}, the basis is \{1, \sqrt{d}\}; if d \equiv 1 \pmod{4}, the basis is \left\{1, \frac{1 + \sqrt{d}}{2}\right\}. Thus, elements of \mathcal{O}_K can be expressed as a + b \sqrt{d} with a, b \in \mathbb{Z} when d \equiv 2 or $3 \pmod{4}, or as a + b \frac{1 + \sqrt{d}}{2} with a, b \in \mathbb{Z} when d \equiv 1 \pmod{4}. Every nonzero element \alpha \in K satisfies a minimal polynomial over \mathbb{Q} of degree at most 2, given by the characteristic equation x^2 - \operatorname{tr}(\alpha) x + N(\alpha) = 0, where \operatorname{tr}(\alpha) is the and N(\alpha) is the of \alpha relative to K/\mathbb{Q}. For \alpha to be a quadratic integer, this minimal polynomial must be monic with integer coefficients, which occurs precisely when \operatorname{tr}(\alpha) and N(\alpha) are integers. For an element \alpha = a + b \sqrt{d} with a, b \in \mathbb{Q}, the trace is $2a and the norm is a^2 - d b^2, so \alpha lies in \mathcal{O}_K if and only if these quantities are integers (adjusting for the basis when d \equiv 1 \pmod{4}). \begin{equation} x^2 - (2a)x + (a^2 - d b^2) = 0 \end{equation}

Basic examples

Quadratic integers provide concrete illustrations of algebraic integers residing in quadratic fields \mathbb{Q}(\sqrt{d}), where d is a square-free integer. A basic example is the imaginary unit i = \sqrt{-1}, which belongs to the quadratic field \mathbb{Q}(\sqrt{-1}) and satisfies the monic polynomial equation x^2 + 1 = 0 with integer coefficients, confirming it as an algebraic integer. Another simple real example is $1 + \sqrt{2} in \mathbb{Q}(\sqrt{2}), whose minimal polynomial is x^2 - 2x - 1 = 0, again monic with integer coefficients. Similarly, the golden ratio \phi = \frac{1 + \sqrt{5}}{2} in \mathbb{Q}(\sqrt{5}) satisfies x^2 - x - 1 = 0, establishing it as a quadratic integer. These examples highlight the standard bases for representing elements in quadratic integer rings. For the imaginary quadratic field with d = -1, the ring consists of Gaussian integers of the form a + bi where a, b \in \mathbb{Z} and i = \sqrt{-1}. In the real quadratic field \mathbb{Q}(\sqrt{2}) where d = 2 \equiv 2 \pmod{4}, elements are expressed as a + b\sqrt{2} with a, b \in \mathbb{Z}. For d = 5 \equiv 1 \pmod{4}, the ring uses the basis \{1, \frac{1 + \sqrt{5}}{2}\}, so quadratic integers take the form a + b \cdot \frac{1 + \sqrt{5}}{2} with a, b \in \mathbb{Z}. Each such element is an because it is a of a monic polynomial with coefficients, as required by the definition in quadratic fields. Quadratic integers play a key role in , particularly in solving Diophantine equations such as x^2 - dy^2 = 1, where solutions correspond to units in the associated real quadratic integer rings.

Representation and Arithmetic

Explicit representation

Quadratic integers in the of a quadratic number field \mathbb{Q}(\sqrt{d}), where d is a not equal to 0 or , can be explicitly represented in a basis over the rational integers \mathbb{Z}. Elements take the form a + b \omega, where a, b \in \mathbb{Z} and \omega is a basis element depending on the congruence class of d modulo 4. The classification arises from the discriminant of the field, which determines the precise structure of the . If d \equiv 2 \pmod{4} or d \equiv 3 \pmod{4}, then \omega = \sqrt{d} and the ring is \mathbb{Z}[\sqrt{d}], consisting of elements a + b \sqrt{d} with a, b \in \mathbb{Z}. If d \equiv 1 \pmod{4}, then \omega = \frac{1 + \sqrt{d}}{2} and the ring is \mathbb{Z}\left[\frac{1 + \sqrt{d}}{2}\right], consisting of elements a + b \frac{1 + \sqrt{d}}{2} with a, b \in \mathbb{Z}. In the latter case, elements can equivalently be written as \frac{m + n \sqrt{d}}{2} where m, n \in \mathbb{Z} are both even or both odd, ensuring integrality. These representations ensure under and , forming a of the algebraic integers. is componentwise: (a + b \omega) + (c + e \omega) = (a + c) + (b + e) \omega. For in the case d \equiv 2, 3 \pmod{4}, (a + b \sqrt{d})(c + e \sqrt{d}) = (a c + b e d) + (a e + b c) \sqrt{d}, with a c + b e d, a e + b c \in \mathbb{Z}. In the case d \equiv 1 \pmod{4}, \left(a + b \frac{1 + \sqrt{d}}{2}\right)\left(c + e \frac{1 + \sqrt{d}}{2}\right) = \left(a c + b e \frac{d - 1}{4}\right) + (a e + b c + b e) \frac{1 + \sqrt{d}}{2}, again yielding integer coefficients. The discriminant \Delta of \mathbb{Q}(\sqrt{d}) is given by \Delta = 4d if d \equiv 2, 3 \pmod{4} and \Delta = d if d \equiv 1 \pmod{4}, which directly governs the choice of basis and ring structure as above. This is the determinant of the trace form on the field and distinguishes the two cases by ensuring the ring consists precisely of elements algebraic in the field.

Norm and conjugation

In quadratic fields K = \mathbb{Q}(\sqrt{d}), where d is a not equal to 0 or 1, the conjugation operation is the non-trivial Galois \sigma: K \to K that fixes \mathbb{Q} pointwise and sends \sqrt{d} to -\sqrt{d}. For an element \alpha = a + b \sqrt{d} with a, b \in \mathbb{Q}, the conjugate is thus \sigma(\alpha) = a - b \sqrt{d}. This map is an , satisfying \sigma^2 = \mathrm{id}, and extends naturally to the \mathcal{O}_K of K. The norm function N_{K/\mathbb{Q}}: K \to \mathbb{Q} is defined as the product N(\alpha) = \alpha \cdot \sigma(\alpha). For \alpha = a + b \sqrt{d}, this yields the explicit formula N(\alpha) = a^2 - d b^2. To derive this from the minimal polynomial, note that \alpha satisfies its characteristic polynomial over \mathbb{Q}, which for a quadratic extension is X^2 - \mathrm{Tr}(\alpha) X + N(\alpha) = 0, where \mathrm{Tr}(\alpha) = \alpha + \sigma(\alpha) = 2a. Substituting gives the monic polynomial X^2 - 2a X + (a^2 - d b^2) = 0, confirming N(\alpha) = a^2 - d b^2. When \alpha is a quadratic integer (i.e., \alpha \in \mathcal{O}_K), N(\alpha) \in \mathbb{Z}, as its minimal polynomial is monic with integer coefficients. The is multiplicative: N(\alpha \beta) = N(\alpha) N(\beta) for all \alpha, \beta \in K, a direct consequence of the homomorphism property of [\sigma](/page/Sigma). It provides a measure of size via |N(\alpha)|, which is preserved under units and plays a role in bounding elements in \mathcal{O}_K. For units u \in \mathcal{O}_K^\times, the norm satisfies N(u) = \pm 1; conversely, if N(\alpha) = \pm 1 for \alpha \in \mathcal{O}_K, then \alpha is invertible with \pm \sigma(\alpha), establishing the units as precisely those elements of \pm 1. This connection aids in studying divisibility, as the norm detects invertibility and factors uniquely in principal ideal domains among quadratic integer rings.

Quadratic Integer Rings

Units

In the \mathcal{O}_K of a quadratic number field K = \mathbb{Q}(\sqrt{d}), where d is a , a is an element u \in \mathcal{O}_K that admits a also in \mathcal{O}_K. This condition is equivalent to the N(u) = \pm 1, since the norm is multiplicative and N(uv) = N(u)N(v), so N(u) = \pm 1 ensures u divides 1 in the ring. The structure of the unit group \mathcal{O}_K^\times is determined by Dirichlet's unit theorem, which states that for quadratic fields, the rank of the unit group is 0 when d < 0 (imaginary quadratic) and 1 when d > 0 (real quadratic). For imaginary quadratic fields, \mathcal{O}_K^\times is a finite cyclic group, typically \{\pm 1\} of order 2, except in the cases d = -1 (order 4) and d = -3 (order 6). For real quadratic fields, \mathcal{O}_K^\times is infinite and isomorphic to \mathbb{Z} \times \{\pm 1\}, generated by -1 and a fundamental unit \varepsilon > 1 of infinite order. To compute the fundamental unit in real quadratic rings, one solves the Pell-like equation x^2 - d y^2 = \pm 1 for the minimal positive solution (x, y) with x > 1, yielding \varepsilon = x + y \sqrt{d}. For example, when d = 2, the solution to x^2 - 2 y^2 = -1 gives the fundamental unit $1 + \sqrt{2}. In imaginary rings, the finite nature simplifies computation to checking elements with small norms.

Examples of complex quadratic integer rings

Complex quadratic integer rings arise in imaginary quadratic fields \mathbb{Q}(\sqrt{d}) where d < 0 is square-free, and their rings of integers exhibit unique properties such as finite unit groups and, in specific cases, unique factorization. These rings are particularly notable for their role in number theory, including factorization and representation problems. The Gaussian integers form the ring of integers for d = -1, denoted \mathbb{Z} where i = \sqrt{-1}. This ring is a Euclidean domain with respect to the norm N(a + bi) = a^2 + b^2, allowing a division algorithm that implies it is a principal ideal domain (PID) and thus has unique factorization. The units in \mathbb{Z} are \{\pm 1, \pm i\}, corresponding to elements of norm 1. Gaussian integers underpin Fermat's theorem on sums of two squares, which states that an odd prime p can be expressed as p = a^2 + b^2 if and only if p \equiv 1 \pmod{4}, proved via factorization in this ring. For d = -3, the Eisenstein integers are the ring \mathbb{Z}[\omega] where \omega = \frac{-1 + \sqrt{-3}}{2}, a primitive cube root of unity satisfying \omega^2 + \omega + 1 = 0. This ring is also Euclidean with norm N(a + b\omega) = a^2 - ab + b^2, making it a PID with unique factorization. The units are \{\pm 1, \pm \omega, \pm \omega^2\}, again the elements of norm 1. Eisenstein integers find applications in the theory of ternary quadratic forms, facilitating proofs of representation theorems for integers by forms like x^2 + y^2 + z^2 - xy - xz - yz. Beyond these, there are exactly nine imaginary quadratic fields where the ring of integers is a PID, hence having unique factorization: for d = -1, -2, -3, -7, -11, -19, -43, -67, -163. In contrast, for d = -5, the ring is \mathbb{Z}[\sqrt{-5}], which is not Euclidean and has class number 2, meaning the ideal class group is non-trivial and unique factorization fails (e.g., $6 = 2 \cdot 3 = (1 + \sqrt{-5})(1 - \sqrt{-5}) with distinct irreducible factorizations up to units). A defining feature of all complex quadratic integer rings is their finite unit group: for most d < 0, it is \{\pm 1\}, except for the cases d = -1 (order 4) and d = -3 (order 6), reflecting the bounded nature of units in imaginary quadratic fields.

Examples of real quadratic integer rings

Real quadratic integer rings arise in quadratic fields \mathbb{Q}(\sqrt{d}) where d > 0 is a square-free positive , featuring infinitely many units due to the two real embeddings. These rings contrast with their counterparts by having unit groups of , generated by -1 and a fundamental \varepsilon > 1, leading to solutions of Pell-like equations x^2 - d y^2 = \pm 1. The , defined as \log \varepsilon, quantifies the of these units under the logarithmic into \mathbb{R}. A prominent example is the \mathbb{Z}[\sqrt{2}] of \mathbb{Q}(\sqrt{2}), where the units are \{\pm (1 + \sqrt{2})^n \mid n \in \mathbb{Z}\}, with $1 + \sqrt{2} as the fundamental unit satisfying x^2 - 2y^2 = -1. Powers of this unit also solve the positive Pell equation x^2 - 2y^2 = 1, such as (3 + 2\sqrt{2}) for n=2. This is with respect to the . Another key example is \mathbb{Z}\left[\frac{1 + \sqrt{5}}{2}\right], of \mathbb{Q}(\sqrt{5}), where the fundamental is the \phi = \frac{1 + \sqrt{5}}{2}, and the units are \{\pm \phi^n \mid n \in \mathbb{Z}\}. This connects to the via Binet's F_n = \frac{\phi^n - (-\phi)^{-n}}{\sqrt{5}}, linking algebraic units to integer sequences. Like \mathbb{Z}[\sqrt{2}], it is . Not all real quadratic integer rings are principal ideal domains; for instance, the ring of \mathbb{Q}(\sqrt{10}) has class number 2, generated by the non-principal above 2, which satisfies \mathfrak{p}^2 = (2) but cannot be generated by a single element of 2. In total, exactly 16 real quadratic integer rings are known to be with respect to the : those for d = 2, 3, 5, 6, 7, 11, 13, 17, 19, 21, 29, 33, 37, 41, 57, 73. Under the two real embeddings \theta_1, \theta_2: K \to \mathbb{R}, the units map to discrete points on the hyperbolas \{(x,y) \in \mathbb{R}^2 : xy = \pm 1\}, with their spacing determined by the .

Principal rings of quadratic integers

A quadratic integer ring, being the ring of integers of a quadratic number field, is a . Such a ring is a if and only if its is trivial, meaning the class number h = 1. This condition is equivalent to the ring possessing unique into prime elements up to units. In the imaginary quadratic case, where the is negative, there are exactly nine quadratic integer rings that are domains. These correspond to the square-free integers d = -1, -2, -3, -7, -11, -19, -43, -67, -163, for which the of \mathbb{Q}(\sqrt{d}) has class number one. This complete was established by the Heegner–Baker–Stark theorem. For real quadratic fields, where the is positive, it is conjectured that infinitely many have class number one and thus are domains, but no proof of this exists, and only finitely many have been verified unconditionally up to sufficiently large discriminants. All real quadratic integer rings with class number one are domains by the general criterion for Dedekind domains.

Euclidean rings of quadratic integers

A quadratic integer ring \mathcal{O}_K of a quadratic number field K = \mathbb{Q}(\sqrt{d}), where d is a not equal to 0 or 1, is called if there exists a \phi: \mathcal{O}_K \setminus \{0\} \to \mathbb{Z}_{\geq 0} such that for every \alpha, \beta \in \mathcal{O}_K with \beta \neq 0, there exist q, \rho \in \mathcal{O}_K satisfying \alpha = \beta q + \rho and either \rho = 0 or \phi(\rho) < \phi(\beta). This property enables the Euclidean algorithm for computing greatest common divisors in the , implying that \mathcal{O}_K is a principal ideal domain. The most commonly used Euclidean function is the absolute value of the field norm N_{K/\mathbb{Q}}, denoted \phi(\alpha) = |N(\alpha)|, where N(\alpha \beta) = N(\alpha) N(\beta) for \alpha, \beta \in K. With this function, the division algorithm requires that for any \alpha \in K and \beta \in \mathcal{O}_K \setminus \{0\}, there exist q, \rho \in \mathcal{O}_K such that \alpha = \beta q + \rho and |N(\rho)| < |N(\beta)|. For imaginary quadratic fields (d < 0), the classification is complete: the rings \mathcal{O}_K are Euclidean if and only if d \in \{-1, -2, -3, -7, -11\}, and in each case, they are Euclidean with respect to the norm function. These five rings—known as the Gaussian integers (d=-1), Eisenstein integers (d=-3), and others—are the only ones among the nine imaginary quadratic fields with class number 1 that admit a Euclidean algorithm. To verify norm-Euclideanity, one checks that |N(\alpha)| < 1 implies \alpha = 0 in K (which holds since norms of nonzero algebraic integers are at least 1 in absolute value), and that for every \xi \in K, there exists \gamma \in \mathcal{O}_K such that |N(\xi - \gamma)| < 1. This condition ensures the remainder in the division algorithm has smaller norm, and it has been exhaustively confirmed for these discriminants through direct computation of "universal side divisors." For real quadratic fields (d > 0), the situation is more complex, as the classification of all rings remains open, though significant progress has been made on norm-Euclidean cases and beyond. The norm-Euclidean real quadratic integer rings are precisely those with d \in \{2, 3, 5, 6, 7, 11, 13, 17, 19, 21, 29, 33, 37, 41, 57, 73\}, comprising 16 fields identified through systematic checks up to large discriminants. Proofs for these follow the same criterion as in the imaginary case: verifying that the ring covers the field adequately so that remainders satisfy |N(\rho)| < |N(\beta)|, often by bounding the minimal norms of differences \xi - \gamma and confirming no larger universal side divisors exist. Additionally, two real quadratic rings are known to be Euclidean with respect to other functions (not the norm): for d = 14, 69. For example, in \mathbb{Q}(\sqrt{69}), a modified function involving the fundamental unit ensures the division property, despite failing the norm condition. It is conjectured that these 18 real quadratic Euclidean rings exhaust the list, with no others existing, though this remains unproven; under the generalized Riemann hypothesis, all real quadratic fields of class number 1 would be Euclidean.

Historical and Modern Developments

Historical origins

The origins of quadratic integers can be traced to ancient Indian mathematics, where solutions to Diophantine equations involving square roots laid the groundwork for their arithmetic properties. In 628 CE, Brahmagupta, in his treatise Brahmasphuṭasiddhānta, introduced the method of composition (samāsa) to solve equations of the form x^2 - d y^2 = k, particularly focusing on Pell equations x^2 - d y^2 = 1 for small values of k. His key contribution, known as Brahmagupta's identity or composition law, states that if (x_1, y_1; m_1) and (x_2, y_2; m_2) satisfy d x_j^2 + m_j = y_j^2, then a composed solution (x_3, y_3; m_3) is given by x_3 = x_1 y_2 + x_2 y_1, y_3 = d x_1 x_2 + y_1 y_2, and m_3 = m_1 m_2; this generates larger solutions from initial ones, implicitly exploiting the multiplicative structure of units in rings such as \mathbb{Z}[\sqrt{d}]. This framework was refined by Bhāskara II in the 12th century through his , where he developed the (cyclic) method, an iterative algorithm that systematically produces the fundamental solution to x^2 - d y^2 = 1 for any positive non-square integer d by selecting auxiliary parameters and composing solutions. Bhāskara's approach built directly on Brahmagupta's composition, enabling solutions to challenging cases like d = 61, and further highlighted the role of units in \mathbb{Z}[\sqrt{d}] as generators of infinite solution families. European developments began in the 17th century with Pierre de Fermat's 1640 statement of the theorem on sums of two squares, asserting that an odd prime p can be expressed as p = x^2 + y^2 if and only if p \equiv 1 \pmod{4}; this motivated the study of arithmetic in the imaginary quadratic ring \mathbb{Z}, though Fermat provided no proof. Leonhard Euler, in the 18th century, supplied the first proof using infinite descent in 1749 and extended investigations to imaginary quadratic fields via quadratic forms and representations, including solutions to Pell-like equations such as x^2 - 61 y^2 = 1. The 19th century brought formal structure, influenced by Ernst Kummer's 1840s theory of ideal numbers, which addressed factorization failures in cyclotomic integers and inspired broader algebraic approaches. In 1871, Richard Dedekind formalized the ring of integers in quadratic fields as a supplement to Dirichlet's Vorlesungen über Zahlentheorie, defining it explicitly: for a quadratic field \mathbb{Q}(\sqrt{D}) with square-free D < 0 or D > 0, the ring consists of elements a + b \omega where \omega = \sqrt{D} if D \equiv 2,3 \pmod{4} (discriminant $4D) or \omega = (1 + \sqrt{D})/2 if D \equiv 1 \pmod{4} (discriminant D); he introduced ideals to restore unique factorization. Concurrently, Carl Gustav Jacob Jacobi's 1832 work on theta functions connected representations by quadratic forms to class numbers in quadratic fields, providing analytic tools for computing ideal class groups.

Modern contributions and applications

In the mid-20th century, significant progress was made in resolving Gauss's for imaginary quadratic fields, which seeks to determine the discriminants for which the has class number one. Kurt Heegner provided a proof in that there are exactly nine such fields, using modular functions and Diophantine to construct explicit j-invariants associated with these rings. This result was later rigorously verified and extended by Harold Stark in 1967, who employed analytic methods involving L-functions to confirm the list and rule out additional cases, thereby fully solving the problem. These advancements relied on connections between quadratic integer rings and elliptic curves with complex multiplication, highlighting the role of quadratic integers in deeper arithmetic structures. Computational methods for determining class numbers in quadratic fields have advanced through the evaluation of Dirichlet L-functions, which encode the of these rings via their special values at s=1. Algorithms based on these L-functions allow efficient of class numbers for large discriminants, often using modular symbols or approximations to bound the error. Such techniques have been implemented to verify class numbers up to discriminants exceeding 10^12, providing empirical on the distribution of class numbers in both real and imaginary quadratic fields. Quadratic integers find applications in , particularly through lattice-based schemes. The cryptosystem, originally over polynomial rings in \mathbb{Z}, has been generalized to rings of \mathbb{Z}[\omega], where \omega is a of unity, yielding the ETRU variant; this leverages the Euclidean structure of for efficient key generation and decryption while enhancing resistance to attacks. In , \mathbb{Z} underpin lattice codes for wireless communications, enabling compute-and-forward protocols that achieve near-capacity performance in multi-user channels by decoding integer linear combinations of transmitted signals. These codes exploit the dense packing properties of the for error correction and signal constellation design. In , quadratic forms arising from quadratic integer rings appear in compactifications on manifolds like K3 × T², where attractor mechanisms for black holes correspond to equivalence classes of binary quadratic forms, linking the of supersymmetric black holes to class group invariants of the underlying quadratic fields. Modern algorithms for units in real quadratic integer rings utilize continued fraction expansions of \sqrt{d}, where the fundamental unit is derived from the convergents of the period; this method efficiently solves and scales to large discriminants by limiting the expansion length to O(\sqrt{d} \log d). Software tools like PARI/GP facilitate computations in quadratic rings, including unit group generation, structure, and evaluations, through built-in functions for number field arithmetic. Quadratic integers serve as foundational models in , particularly in , where the Hilbert class field of an imaginary quadratic field is generated by j-invariants of elliptic curves with complex multiplication by the , explicitly constructing abelian extensions via these structures.

References

  1. [1]
    None
    Below is a merged summary of the segments on "Quadratic Integers and Quadratic Fields" from Trifković (2013), consolidating all information into a single, comprehensive response. To maximize detail and clarity, I will use a combination of narrative text and tables where appropriate (e.g., for key properties like norms, units, and ring structures). The response retains all unique details from the provided summaries while avoiding redundancy.
  2. [2]
    [PDF] notes on introductory algebraic number theory - UChicago Math
    Aug 20, 2013 · The prime factorization theorem says that every integer can be factored uniquely (up to sign) into a product of prime numbers; i.e. for all z in ...
  3. [3]
    [PDF] Math 784: algebraic NUMBER THEORY
    An algebraic number α is an algebraic integer if it is a root of some monic polynomial f(x) ∈ Z[x] (i.e., a polynomial f(x) with integer coefficients and ...
  4. [4]
    [PDF] Algebraic Number Theory - James Milne
    the ring of integers in the number field, the ideals and units in the ring of.
  5. [5]
    [PDF] Algebraic Number Theory Ben Green - People
    4 Algebraic Number Theory at. Oxford in Hilary Term 2020. They cover the examinable material of that course as well as some extra material on the link ...
  6. [6]
    [PDF] math3704: algebraic number theory - metaphor
    (the “Golden Ratio”) is an algebraic integer, since it satisfies τ2 − τ − 1 = 0. We will later determine all the algebraic integers in quadratic fields. Clearly ...
  7. [7]
    [PDF] Gaussian Integers and Other Quadratic Integer Rings - DiVA portal
    Taking the minimal polynomial of α we get x2 - 2ax + (a2 - b2D). Since the ... = ϕ is known as the golden ratio. An element a+bϕ belongs to the real ...
  8. [8]
    [PDF] Math 210B. Quadratic integer rings 1. Computing the integral ...
    Already with quadratic integer rings one can begin to see some ring-theoretic subtleties emerge. As a basic example, one might wonder: for a finite ...
  9. [9]
    [PDF] Pell's Equation and Fundamental Units
    The fundamental unit of the ring of algebraic integers in a real quadratic number field is a generator of the group of units (mod ±1). For the subring Z ...
  10. [10]
    [PDF] Contents 8 Quadratic Integer Rings - Evan Dummit
    The question of when primes split, remain inert, or ramify is a fundamental object of study in algebraic number theory. 27. Page 28. 8.3 Applications of ...
  11. [11]
    [PDF] dirichlet's unit theorem - keith conrad
    Introduction. Dirichlet's unit theorem describes the structure of the unit group of orders in a number field. Theorem 1.1 (Dirichlet, 1846).
  12. [12]
    [PDF] Fermat's Theorem on Sums of Squares - Williams College
    One can show that the norm serves as an appropriate Euclidean norm; the Gaussian integers are a Euclidean Domain. One can show that 𝑁 𝛼 = 1 if and only if 𝛼 is ...
  13. [13]
    [PDF] The Eisenstein integers and cubic reciprocity - Uppsala University
    Proposition 3.2. The ring of Eisenstein integers, Z[ω], is a sub-ring of (C,+,·). (a + bω)+(c + dω)=(a + c)+(b + d)ω (a + bω)(c + dω) = ac + adω + bcω + bdω2 = ...
  14. [14]
    [PDF] An Exposition of the Eisenstein Integers - Eastern Illinois University
    May 1, 2016 · We also provide the necessary background to understand how the imaginary ring of quadratic integers behaves. An example of said ring are complex ...
  15. [15]
    [PDF] Gauss' Class Number Problems for Imaginary Quadratic Fields
    Abstract. This paper examines Gauss' class number problems for imaginary quadratic fields, with particular emphasis on the class number one problem. We.
  16. [16]
    [PDF] The ideal class number formula for an imaginary quadratic field
    The more complicated unit group struc- ture for real quadratic fields is one reason that the class number formula is easier in the imaginary case. 3. The ideals ...
  17. [17]
    [PDF] the structure of unit groups - UChicago Math
    Aug 29, 2014 · Now, consider the minimal polynomial mα(x) = xn + an−1xn−1 + ... integer and the aforementioned polynomial is a power of the minimal polynomial.<|separator|>
  18. [18]
    [PDF] Euclidean rings of S-integers in complex quadratic fields - arXiv
    Mar 29, 2022 · ... [2]). It turns out that K = Q(. √ d) with d > 0 squarefree is norm-Euclidean if and only if d = 2,3,5,6,7,11,13,17,19,21,29,33,37,41,57,73 ...
  19. [19]
    [PDF] Chapter 15 Q( √ 5) and the golden ratio
    Recall that the Fibonacci sequence consists of the numbers. 0,1,1,2,3,5,8,13,... defined by the linear recurrence relation. Fn+1 = Fn + Fn−1, with initial ...<|separator|>
  20. [20]
    [PDF] The Class Number Formula for Quadratic Fields and Related Results
    Jan 31, 2016 · Since p is not principal but p2 = (2) is principal, p has order 2 and hence the ideal class group consists of two ideal classes. Therefore ...
  21. [21]
    on complex quadratic fields with class number equal to one(1) - jstor
    Received by the editors May 6, 1965. (1) This paper represents a portion of my Ph. D. dissertation written under the supervision of Professor D. H. Lehmer.
  22. [22]
    [PDF] Class Numbers of Quadratic Fields
    1. In the case of imaginary quadratic fields, Cohen and Lenstra predict the following: Conjecture 4. (Cohen-Lenstra) Let K be ...
  23. [23]
    [PDF] DIPLOMARBEIT Quadratic Number Fields that are Euclidean but not ...
    All Norm-Euclidean quadratic number fields Q(. √ d) for squarefree d 6= 0,1 are known since 1950: Chatland & Davenport [3] and independently Inkeri [12] ...
  24. [24]
    [PDF] Math 71: Principal Ideal Domains, Quadratic Integer Rings, and ...
    Nov 7, 2023 · This leads us to consider the absolute value of the field norm. Definition 3.4. The quadratic integer ring OD ⊂ Q(. √D) is norm-Euclidean if it ...
  25. [25]
    quadratic imaginary norm-Euclidean number fields - PlanetMath.org
    Mar 22, 2013 · Now |d|≥15 | d | ≥ 15 . The integers of Q(√d) ℚ ⁢ ( d ) have the form ϰ=a+b√d2 ϰ = a + b ⁢ d 2 with 2|a−b 2 | a - b . Suppose that γ=14+14√d ...
  26. [26]
    [PDF] Integers in quadratic fields; EDs and PIDs.
    Feb 9, 2012 · We call any such f a Euclidean function on R. We say that a number field is Euclidean if its ring of integers R is an ED. We say it's norm-.
  27. [27]
    [PDF] Admissible primes and Euclidean quadratic fields
    It is known that Q(. √ d), d > 0 is norm-Euclidean if and only if d = 2, 3, 5, 6, 7, 11, 13, 17,. 19, 21, 29, 33, 37, 41, 57, 73. Thus, we have a complete list ...
  28. [28]
    Euclidean real quadratic fields - MathOverflow
    Oct 17, 2013 · I want to know if there are any approaches in showing that there are infinitely many Euclidean real quadratic fields by explicitly constructing the Euclidean ...A question about non-norm-euclidean real quadratic fieldsAbout list of discriminants of real quadratic fields with narrow class ...More results from mathoverflow.net
  29. [29]
    ℤ[14] is Euclidean - ResearchGate
    Aug 6, 2025 · In this chapter we present examples of norm-Euclidean quadratic number fields and apply the results to the Fermat equations with exponents 3, 4, ...
  30. [30]
    Pell's equation - MacTutor History of Mathematics
    So Brahmagupta was able to show that if he could find ( a , b ) (a, b) (a,b) which "nearly" satisfied Pell's equation in the sense that n a 2 + k = b 2 na^{2} + ...
  31. [31]
    [PDF] On the Brahmagupta- Fermat-Pell Equation: The Chakrav¯ala ... - HAL
    Jul 29, 2023 · In the following pages we take a fresh look at the ancient Indian Chakrav¯ala or Cyclic algorithm for solving the Brahmagupta-Fermat-Pell ...
  32. [32]
    [PDF] Gaussian Integers and Dedekind's Creation of an Ideal
    The first published proof of the Sum of Four Squares Theorem was given in 1770 by the French mathematician J. L. Lagrange (1736-. 1813). Before then, Fermat ...
  33. [33]
    [PDF] Dedekind's 1871 version of the theory of ideals∗ - andrew.cmu.ed
    Mar 19, 2004 · This was to remedy the situation by introducing, for each such ring of integers, an enlarged domain of divisors, and showing that each.
  34. [34]
    (PDF) Jacobi and Kummer's ideal numbers - ResearchGate
    Aug 6, 2025 · In this article we give a modern interpretation of Kummer's ideal numbers and show how they developed from Jacobi's work on cyclotomy, ...<|separator|>
  35. [35]
    Diophantische Analysis und Modulfunktionen - SpringerLink
    Diophantische Analysis und Modulfunktionen. Download PDF · Download PDF. Published: September 1952. Diophantische Analysis und Modulfunktionen. Kurt Heegner.
  36. [36]
    A complete determination of the complex quadratic fields of class ...
    April 1967 A complete determination of the complex quadratic fields of class-number one. H. M. Stark · DOWNLOAD PDF + SAVE TO MY LIBRARY. Michigan Math.
  37. [37]
    [PDF] Class Field Theory
    Class field theory describes the abelian extensions of a local or global field in terms of the arithmetic of the field itself.
  38. [38]
  39. [39]
  40. [40]
    Catalogue of GP/PARI Functions: General number fields
    Number field structures. Let K = ℚ[X] / (T) a number field, ℤK its ring of integers, T ∈ ℤ[X] is monic. Three basic number field structures can be attached ...
  41. [41]
    [PDF] class field theory for number fields and complex multiplication
    We state the main results of class field theory for a general number field, and then specialize to the case where K is imaginary quadratic. By look- ing at ...