Fact-checked by Grok 2 weeks ago

Primitive element theorem

The primitive element theorem is a fundamental result in field theory stating that if E/F is a finite of fields, then there exists an \theta \in E such that E = F(\theta), meaning the extension is and generated by a single over the base field F. This theorem also provides an equivalent condition: a finite extension E/F admits a primitive if and only if there are only finitely many intermediate fields between F and E. In greater detail, separability ensures that the minimal polynomials of the generating elements over F have distinct roots, allowing the construction of such a \theta explicitly, often as a \theta = \alpha + c\beta where \alpha, \beta \in E generate E over F and c \in F is chosen to avoid certain "bad" values that would prevent . For extensions of characteristic zero or finite s, all finite extensions are separable, so the theorem applies universally in these cases. The proof typically proceeds by induction on the degree of the extension, first handling the two-element case using the infinitude of the base to select an appropriate c, and then reducing larger generating sets iteratively. The theorem's significance lies in simplifying the algebraic structure of extensions, enabling representations of complex fields like number fields or function fields as simple adjunctions, which facilitates computations in and . It has applications beyond , including in where it aids in analyzing algebras over separable extensions by reducing to simple cases. In characteristic p > 0, purely inseparable extensions may lack primitive elements, highlighting the necessity of the separability condition.

Terminology

Primitive Elements

In field theory, an element \alpha in a E/F is called a primitive element if E = F(\alpha), meaning that E is generated by \alpha over the base field F. This implies that every element of E can be expressed uniquely as a in \alpha with coefficients in F. For a finite extension E/F of degree n, if \alpha is a primitive element, then the set \{1, \alpha, \alpha^2, \dots, \alpha^{n-1}\} forms a basis for E as a over F, known as a power basis. Moreover, the minimal polynomial of \alpha over F must be irreducible and of degree exactly n, which equals the degree of the extension. While elements exist for every finite extension under certain conditions, they may not exist in infinite extensions, where the extension cannot always be generated by a single element; however, the concept is primarily relevant to finite cases. Such extensions generated by a single element are termed extensions.

Simple and Separable Extensions

A E/F is called a if there exists an \alpha \in E such that E = F(\alpha). Such an \alpha is known as a primitive element for the extension. A E/F is separable if every \alpha \in E is separable over F, meaning that the minimal polynomial of \alpha over F has distinct roots in its . Equivalently, for a finite extension E/F, separability holds if the number of F-homomorphisms from E into an of F equals the degree [E:F]. This embedding highlights the "full" preserved in separable extensions. All finite separable extensions are simple, as established by the primitive element theorem. In contrast, infinite extensions, such as the algebraic closure of a field like \mathbb{Q}, have infinite degree and thus cannot be simple algebraic extensions. These concepts classify when a primitive element exists, extending the basic notion of primitive elements to broader extension properties.

Examples

Over the Rationals

A classic example illustrating the primitive element theorem over the rationals is the extension \mathbb{Q}(\sqrt{2}, \sqrt{3})/\mathbb{Q}, which has degree 4 since the minimal polynomials x^2 - 2 and x^2 - 3 are irreducible and the extensions are linearly disjoint. The element \alpha = \sqrt{2} + \sqrt{3} generates this extension, as \mathbb{Q}(\alpha) = \mathbb{Q}(\sqrt{2}, \sqrt{3}), with the minimal polynomial of \alpha over \mathbb{Q} given by x^4 - 10x^2 + 1 = 0. To express the adjoined elements in terms of \alpha, first compute \alpha^2 = 5 + 2\sqrt{6}, so \sqrt{6} = (\alpha^2 - 5)/2. Then, \alpha \sqrt{6} = 3\sqrt{2} + 2\sqrt{3}. Solving the linear system \sqrt{2} + \sqrt{3} = \alpha and $3\sqrt{2} + 2\sqrt{3} = \alpha \sqrt{6} yields \sqrt{2} = (\alpha^3 - 9\alpha)/2 and \sqrt{3} = (-\alpha^3 + 11\alpha)/2. This confirms that both \sqrt{2} and \sqrt{3} lie in \mathbb{Q}(\alpha), establishing \alpha as a primitive element via direct computation. The extension \mathbb{Q}(\sqrt{2}, \sqrt{3})/\mathbb{Q} is Galois with isomorphic to the \mathbb{Z}/2\mathbb{Z} \times \mathbb{Z}/2\mathbb{Z}, generated by the automorphisms sending \sqrt{2} \mapsto -\sqrt{2} (fixing \sqrt{3}) and \sqrt{3} \mapsto -\sqrt{3} (fixing \sqrt{2}). These automorphisms extend to \mathbb{Q}(\alpha) by mapping \alpha to \pm \sqrt{2} \pm \sqrt{3} (all sign combinations), and since the fixed field of the full group is \mathbb{Q}, the degree matches, verifying that \alpha generates the entire extension. Another example involves cyclotomic fields, as in the extension \mathbb{Q}(\sqrt{{grok:render&&&type=render_inline_citation&&&citation_id=3&&&citation_type=wikipedia}}{2}, \omega)/\mathbb{Q}, where \omega is a primitive cube root of unity satisfying \omega^2 + \omega + 1 = 0. This is the splitting field of x^3 - 2 over \mathbb{Q}, with degree 6, as [\mathbb{Q}(\sqrt{{grok:render&&&type=render_inline_citation&&&citation_id=3&&&citation_type=wikipedia}}{2}):\mathbb{Q}] = 3 and [\mathbb{Q}(\omega):\mathbb{Q}] = 2, and the polynomials remain irreducible over the intermediate fields. By the primitive element theorem, since the extension is separable (characteristic zero), it is simple; an explicit primitive element is \beta = \sqrt{{grok:render&&&type=render_inline_citation&&&citation_id=3&&&citation_type=wikipedia}}{2} + \omega, whose minimal polynomial over \mathbb{Q} has degree 6, highlighting the role of cyclotomic elements in generating such radical extensions.

In Finite Fields

In finite fields, every extension of a finite field by another finite field is , as guaranteed by the primitive element theorem under the separability condition, which holds universally in this setting since finite fields are perfect. Specifically, for the extension \mathrm{GF}(p^n)/\mathrm{GF}(p) where p is prime and n \geq 1, the degree [\mathrm{GF}(p^n):\mathrm{GF}(p)] = n, and the larger field is generated by a single primitive element \alpha \in \mathrm{GF}(p^n) such that \{\alpha^k : k = 0, 1, \dots, p^n - 2\} generates the \mathrm{GF}(p^n)^\times. This primitive element \alpha satisfies a minimal of degree n over \mathrm{GF}(p), providing a basis \{1, \alpha, \alpha^2, \dots, \alpha^{n-1}\} for \mathrm{GF}(p^n) as a over \mathrm{GF}(p). The key property enabling this is that the of any \mathrm{GF}(q) with q = p^n elements is cyclic of order q-1. Consequently, any of \mathrm{GF}(q)^\times—termed a primitive element in this context—also generates the entire as a over the prime subfield \mathrm{GF}(p), aligning the algebraic and group-theoretic notions of primitivity. There are exactly \phi(q-1) such primitive elements, where \phi is , comprising a positive proportion of the nonzero elements as q grows. A concrete example is the extension \mathrm{GF}(4)/\mathrm{GF}(2), which has degree $2. This field is constructed as \mathrm{GF}(2)/(x^2 + x + 1), where \alphais a root of the [irreducible polynomial](/page/Irreducible_polynomial)x^2 + x + 1 = 0over\mathrm{GF}(2).[18] The elements of \mathrm{GF}(4)are{0, 1, \alpha, \alpha + 1 = \alpha^2}, with the nonzero elements forming the cyclic [multiplicative group](/page/Multiplicative_group) \langle \alpha \rangle = {1, \alpha, \alpha^2} of order $3, confirming \alpha as a primitive element. The basis over \mathrm{GF}(2) is \{1, \alpha\}.

The Theorem

Statement

The primitive element theorem asserts that if E/F is a finite separable , then there exists an element \alpha \in E, called a primitive element, such that E = F(\alpha). In particular, when the characteristic of F is zero, every is separable, so the theorem implies that every finite extension of a characteristic zero , such as \mathbb{Q}, is . Within , for a separable extension E/F, the existence of a primitive element \alpha \in E ensures that E is the of the irreducible minimal of \alpha over F, and the \mathrm{Gal}(E/F) acts transitively on the roots of this polynomial.

Assumptions and Scope

The primitive element theorem requires that the field extension L/K be both finite and separable. A finite extension means [L:K] < \infty, ensuring the extension has finite degree as a vector space over K. Separability demands that every element of L is separable over K, meaning its minimal polynomial over K has distinct roots in an algebraic closure; equivalently, the extension has a separating transcendence basis or the number of K-homomorphisms from L to an algebraic closure equals the degree [L:K]. The finiteness assumption is essential because any simple algebraic extension K(\alpha)/K has finite degree equal to the degree of the minimal polynomial of \alpha over K. Consequently, infinite algebraic extensions cannot be simple; for example, the extension of the rationals \mathbb{Q} obtained by adjoining the square roots of all prime numbers is infinite and possesses infinitely many intermediate subfields, precluding a primitive element. The theorem's scope encompasses algebraic number fields, where the characteristic-zero base field \mathbb{Q} guarantees that all finite extensions are separable, allowing every finite extension to be generated by a single algebraic integer or number. It also applies to separable extensions of function fields over algebraically closed constants and to all finite extensions of finite fields, as finite fields are perfect and thus yield only separable extensions. However, the theorem does not hold for inseparable extensions, which arise exclusively in positive characteristic when minimal polynomials have multiple roots. This result connects to Steinitz's theorem, which characterizes finite extensions admitting a primitive element precisely as those with only finitely many intermediate subfields. In Galois theory, the primitive element simplifies the correspondence between subfields and subgroups of the Galois group by reducing the extension to one generated by a single element whose conjugates determine the group's action.

Positive Characteristic

Separable Case

In fields of positive characteristic p, the primitive element theorem asserts that every finite separable extension L/K is simple, meaning L = K(\gamma) for some primitive element \gamma \in L, analogous to the situation in characteristic zero. An extension L/K is separable if the minimal polynomial of every \alpha \in L over K has distinct roots in an algebraic closure, ensuring no multiple roots arise from the characteristic p dividing the degree in a way that causes inseparability. A canonical example of such separable extensions in characteristic p is the Artin-Schreier extension, obtained by adjoining a root b to K satisfying the equation x^p - x - a = 0, where a \in K lies outside the image of the Artin-Schreier map \wp: K \to K given by \wp(x) = x^p - x. These extensions are Galois of degree p, hence separable, and are simple by construction with b serving as the primitive element, as the minimal polynomial x^p - x - a is irreducible and separable under the given condition on a. The existence of a primitive element in the separable case is verified through the Galois correspondence: separability guarantees that L/K is Galois with the order of the Galois group equaling the degree [L:K], allowing the standard proof to construct \gamma as a suitable linear combination of basis elements that separates the intermediate fields via the group action. Thus, even in positive characteristic, separability ensures the extension admits a primitive element through the same combinatorial mechanism as in the general separable case.

Inseparable Counterexamples

In fields of positive characteristic p > 0, the primitive element theorem fails for inseparable extensions, as demonstrated by specific counterexamples where the extension has p^2 but cannot be generated by a single element. Consider the extension k(T, U)/k(T^p, U^p), where k is a of characteristic p and T, U are indeterminates; this extension has p^2 over the base k(T^p, U^p), but any single element \alpha in the extension generates a subextension of degree at most p, so no primitive element exists. The inseparability arises because the minimal polynomials of T and U over k(T^p, U^p) are of the form X^p - T^p and Y^p - U^p, which are inseparable, having all roots equal in a (a single root of multiplicity p). Such polynomials factor as (X - a)^p for some a, leading to extensions that are not when composed, as the Frobenius map z \mapsto z^p collapses the and prevents a single generator from capturing the full . A simpler purely inseparable example is the extension k(x)/k(x^p), which has degree p and is simple, generated by x whose minimal polynomial is the inseparable X^p - x^p. However, composites like k(x, y)/k(x^p, y^p) (with x, y indeterminates) extend this failure to higher degrees, again lacking a primitive element. These counterexamples imply that a finite extension admits a primitive element only if its inseparable degree is 1, meaning the extension is separable; otherwise, inseparability introduces irreducible obstructions to simplicity.

Proof

Infinite Base Fields

In a finite separable extension E = F(\alpha_1, \dots, \alpha_n) where the base field F is infinite, the proof of the primitive element theorem proceeds by showing that the set of non-primitive elements in E is "thin" in the sense that it avoids dense subsets, ensuring the existence of primitive elements. Specifically, the non-primitive elements form a proper subvariety of E viewed as an over F, but the argument focuses on one-dimensional slices where primitive elements are cofinite. The proof uses on n, the number of generators. For the base case n=1, E = F(\alpha_1) is already , so \alpha_1 is . Assume the result holds for n-1: let K = F(\alpha_1, \dots, \alpha_{n-1}) = F(\delta) for some \delta \in K. Then E = K(\alpha_n), and since F is , so is K as a finite extension. It remains to show E = K(\eta) for some \eta \in E, which reduces to the case of adjoining one over an . The key argument for adjoining one element relies on perturbation via linear combinations. Consider distinct elements \alpha, \beta \in E with E = F(\alpha, \beta), where \beta is separable over F(\alpha); seek \eta = \alpha + c \beta with c \in F such that E = F(\eta). Let m = [F(\alpha):F] and let the distinct conjugates of \alpha over F be \alpha_1 = \alpha, \dots, \alpha_m, with minimal polynomial f(x) \in F. Let k = [F(\beta):F] and the distinct conjugates of \beta be \beta_1 = \beta, \dots, \beta_k, with minimal polynomial g(x) \in F. The element \eta generates E if [F(\eta):F] = mk, which occurs precisely when the conjugates \alpha_i + c \beta_j (for $1 \leq i \leq m, $1 \leq j \leq k) are all distinct. The values of c that fail this are "bad," arising when \alpha_i + c \beta_j = \alpha_p + c \beta_q for distinct pairs (i,j) \neq (p,q). Assuming \beta_j \neq \beta_q (which holds by separability of \beta), this rearranges to c = (\alpha_p - \alpha_i)/(\beta_j - \beta_q), yielding at most m(k-1) + k(m-1) such rational expressions in the roots, hence finitely many bad c \in F. Since F is infinite, there exist infinitely many good c, so \eta = \alpha + c \beta is primitive for such choices. By induction, this adjoins the nth element to the primitive generator of the first n-1, yielding a primitive element for E/F. This perturbation leverages the \mathrm{Gal}(E/F) (or more generally, the set of F-embeddings of E into an ). The distinct embeddings \sigma: E \to \overline{F} are linearly independent over F as functions on E, meaning that if \sum_{\sigma} a_{\sigma} \sigma(\alpha) = 0 for all \alpha \in E with a_{\sigma} \in F, then all a_{\sigma} = 0. This independence ensures that for a primitive \eta, the images \{\sigma(\eta) \mid \sigma \in \mathrm{Gal}(E/F)\} span E as an F-vector space, confirming [F(\eta):F] = [E:F] and thus simplicity. In the setup, the distinctness of conjugates \sigma(\eta) follows from avoiding the bad c, aligning with this independence to guarantee the full degree. The density argument completes the picture: along the line \{\alpha + c \beta \mid c \in F\} in E, the primitive elements are cofinite (all but finitely many), and since such lines cover dense subsets of E, primitive elements are dense in E. This not only proves existence but shows there are infinitely many primitive elements when F is infinite.

Finite Base Fields

When the base field F is finite with |F| = q, any finite extension E/F of degree n = [E : F] is likewise finite with |E| = q^n, and thus E is a . The E^\times of nonzero elements in E is finite of q^n - 1 and cyclic; this follows from the structure of finite fields, where the equation x^{q^n - 1} - 1 = 0 has exactly q^n - 1 in E, and for each d of q^n - 1, there is a unique cyclic of d. Consequently, E^\times admits primitive elements, meaning generators \gamma \in E^\times of exactly q^n - 1; the number of such generators is \phi(q^n - 1) > 0, where \phi is . To show that such a \gamma generates E as a over F, suppose for contradiction that [F(\gamma) : F] = d < n. Then \gamma lies in some intermediate L with [L : F] = d, so the of \gamma divides |L^\times| = q^d - 1. But the of \gamma is q^n - 1, implying q^n - 1 divides q^d - 1. This divisibility relation holds if and only if n divides d, contradicting d < n. Thus, [F(\gamma) : F] = n, so E = F(\gamma). This direct construction leverages the cyclic structure unique to finite fields and avoids the density arguments required in the infinite base field case; alternatively, one may embed E into its over F (which remains finite) and apply perturbations within F to separate embeddings, but the generator approach suffices for completeness.

History

Early Developments

The ideas underlying the primitive element theorem originated in the late with Joseph-Louis Lagrange's investigations into the solvability of polynomial equations. In his 1771 paper Réflexions sur la résolution algébrique des équations, Lagrange introduced resolvents as auxiliary polynomials whose roots are formed by linear combinations of the original equation's roots, weighted by powers of primitive roots of unity. This method demonstrated that cubic and quartic equations could be reduced to solving lower-degree equations, effectively showing their solvability via a single generating element in the extension field. Preceding Évariste Galois's more systematic approach, explored related concepts in 1799 through his Teoria generale delle equazioni, where he analyzed groups acting on equation and distinguished between primitive and imprimitive groups. Ruffini's work touched on group actions relevant to extensions but did not fully formalize the role of primitive elements, serving instead as a precursor to by emphasizing structures in algebraic solvability. Galois advanced these foundations significantly in his 1831 memoir Mémoire sur les conditions de résolubilité des équations par (published posthumously in 1846), where he implicitly employed primitive elements to construct Galois groups for radical extensions of and proved versions of the for specific separable extensions in characteristic zero, such as those arising from irreducible polynomials of prime degree. In this work, Galois recognized that abelian extensions over \mathbb{Q} are precisely those generated by radicals, thereby connecting primitive elements to the broader framework of solvability.

Modern Formulations

In 1910, Ernst Steinitz established the general form of the primitive element theorem for finite separable field extensions in his seminal work Algebraische Theorie der Körper, demonstrating that any such extension L/K admits a primitive element \alpha \in L such that L = K(\alpha). This proof extended earlier results by incorporating separability as a key condition, ensuring the extension's simplicity even over fields of positive characteristic, provided the extension is separable. During the 1930s, provided a reformulation of the theorem within his modern approach to , leveraging the independence of field automorphisms to simplify proofs. Artin's method treats the extension as a module over the fixed of its , showing that the degree equals the group order and facilitating the construction of a primitive element via linear combinations that avoid fixed points under non-identity automorphisms. This automorphism-based perspective streamlined earlier inductive arguments and integrated seamlessly with the . Post-1910 developments linked the theorem to , particularly in addressing Hilbert's 12th problem on explicit constructions of abelian extensions. The primitive element theorem enables the description of Hilbert class fields as simple extensions K(\alpha) for number fields K, aiding explicit computations of unramified abelian extensions via j-invariants or modular functions. In computational algebra, algorithms for finding primitive elements in number fields emerged prominently, with implementations in systems like that compute absolute fields via primitive elements over \mathbb{Q}. These algorithms, often based on resultant computations or , facilitate practical tasks such as factorization and group calculations in high-degree extensions. Effective bounds on the degree of primitive elements addressed longstanding gaps, with results showing that for a L = K(\alpha_1, \dots, \alpha_n) of d = [L:K], a primitive element \theta = \sum u_i \alpha_i exists with minimal degree at most d, and explicit constructions bound the coefficients u_i by O(d^2) in infinite base fields. Such bounds, refined in the , support algorithmic efficiency in symbolic computation. Applications in cryptography utilized primitive elements over finite fields, where generators of the multiplicative group \mathbb{F}_{q^n}^\times underpin pseudorandom number generation and discrete logarithm-based protocols. For instance, in elliptic curve cryptography over \mathbb{F}_q, primitive elements ensure efficient field representations for scalar multiplication. As of 2025, the theorem remains foundational without major revisions, though its role has expanded in algorithmic algebraic geometry, where primitive elements parameterize étale covers of varieties and support computations in arithmetic geometry software for solving Diophantine equations.

References

  1. [1]
    Lemma 9.19.1 (030N): Primitive element—The Stacks project
    Lemma 9.19.1 (Primitive element). Let E/F be a finite extension of fields. The following are equivalent. there exists a primitive element for E over F, and.
  2. [2]
    [PDF] The Primitive Element Theorem.
    The Primitive Element Theorem. Assume that F and K are subfields of C and that K/F is a finite extension. Then K = F(θ) for some element θ in K.
  3. [3]
    [PDF] Mathematics 6310 The Primitive Element Theorem Ken Brown ...
    Given a field extension K/F, an element α ∈ K is said to be separable over F if it is algebraic over F and its minimal polynomial over F is separable. Recall ...
  4. [4]
    [PDF] primitive element theorem and normal basis theorem - OSU Math
    For a field extension L/K, a primitive element is any algebraic α ∈ L such that L = K(α). A necessary condition for the existence of such an element is that ...
  5. [5]
    On the Theorem of the Primitive Element with Applications to the ...
    Let K/F be a finite separable field extension and let x, y ∈ K. When is F[x, y] = F[αx + βy] for some nonzero elements α, β ∈ F?
  6. [6]
    Section 9.19 (09HZ): Primitive elements—The Stacks project
    Let E/F be a finite extension of fields. An element \alpha \in E is called a primitive element of E over F if E = F(\alpha).
  7. [7]
    [PDF] an introduction to the theory of field extensions - UChicago Math
    over F. Definition 3.18. If the field K is generated by a single element α over F, K = F(α), then K is said to be a simple extension of F and the element α is ...
  8. [8]
    [PDF] SEPARABILITY 1. Introduction Let K be a field. We are going to look ...
    In Section 3 we will define what it means for a field extension to be separable and then prove the primitive element theorem, an important result about ...
  9. [9]
    Algebraic Closure -- from Wolfram MathWorld
    ### Summary on Algebraic Closures as Simple Extensions and Degree Over Base Field
  10. [10]
    [PDF] ALGEBRA HW 8 1 (a): Find the degree of α = √ 2+ √ 3 over Q, and ...
    Hence, f is the minimal polynomial of √ 2 + √ 3 over Q and so deg α = deg f = 4.
  11. [11]
    Primitive Element -- from Wolfram MathWorld
    ... `PrimitiveElement`. For example, a primitive element of Q(sqrt(2),sqrt(3))/Q is given by b=sqrt(2)+sqrt(3) , with. sqrt(2), = 1/2b(b^2-9). (1). sqrt(3), = 1/2b( ...
  12. [12]
    [PDF] Section 6: Field and Galois theory
    The minimal polynomial of r over F is the irreducible polynomial in F[x] of which r is a root. It is unique up to scalar multiplication. Examples. √. 2 has ...
  13. [13]
    [PDF] the splitting field of x3 − 2 over q - Keith Conrad
    If p ≡ 2 mod 3, then 2 is a cube mod p and there is no primitive cube root of unity in. Z/pZ, so (since h(k) = 1). (p)=(x1)(x2), N(x1) = p, N(x2) = p2. Here ...
  14. [14]
    [PDF] The Multiplicative Group of a Finite Field
    The purpose of these notes is to give a proof that the multiplicative group of a finite field is cyclic, without using the classification of finite abelian.
  15. [15]
    [PDF] 3 Finite fields and integer arithmetic - MIT Mathematics
    Feb 13, 2019 · Theorem 3.11. Every finite subgroup of the multiplicative group of a field is cyclic. Proof. Let k be a field, let G be a subgroup of k× of ...
  16. [16]
    [PDF] how to construct them, properties of elements in a finite field, and ...
    If F is a finite field, the group F× is cyclic. Proof. Let q = |F|, so |F×| = q − 1. Let m be the maximal order of the elements ...
  17. [17]
    Primitive Polynomial List - By Arash Partow
    The following is a list of primitive irreducible polynomials for generating elements of a binary extension field GF(2 m ) from a base finite field.
  18. [18]
    [PDF] Lecture 5: Algebra 3: Irreducible, Primitive and Minimal Polynomials
    GF(4) as an extension field of GF(2). – f(X)=X2+X+1 is a primitive polynomial of degree 2 in GF(2). – m = 2. – The root of f(X) in GF(22) is a primitive ...
  19. [19]
    [PDF] An Introduction to Number Theory JJP Veerman
    Apr 30, 2023 · 5, et cetera (square roots of primes) are irrational. (Hint: use ... infinite field is a simple extension. Remark 7.20. Any single ...
  20. [20]
    [PDF] 4 Étale algebras, norm and trace - 4.1 Separability - MIT Mathematics
    Sep 20, 2021 · Definition 4.18. A field K is perfect if every algebraic extension of K is separable. All fields of characteristic zero are perfect. Perfect ...
  21. [21]
    [PDF] Section V.6. Separability
    Feb 14, 2016 · Let F be a field of characteristic p 6= 0. Lemma V.5.5 shows that for ... The Primitive Element Theorem. Let F be a finite dimensional ...
  22. [22]
    Section 9.25 (09I7): Artin-Schreier extensions—The Stacks project
    9.25 Artin-Schreier extensions. Let K be a field of characteristic p > 0. Let a \in K. Let L be an extension of K obtained by adjoining a root b of the ...Missing: primitive | Show results with:primitive
  23. [23]
    [PDF] Fields and Galois Theory - James Milne
    For example, the extension 𝔽𝑝(𝑇) of 𝔽𝑝(𝑇𝑝) is inseparable because 𝑇 has minimal polynomial 𝑋. 𝑝 − 𝑇𝑝 . Definition 3.7 An algebraic extension 𝐸 ...
  24. [24]
    [PDF] Linear independence of characters - Keith Conrad
    Using a normal basis and the trace function, let's see how to write down a primitive element for every intermediate extension in a finite Galois extension.
  25. [25]
  26. [26]
    [PDF] The Structure of Finite Fields - Henry D. Pfister
    In a cyclic group with n elements, there are ϕ(n) ≥ 1 primitive elements. If d | n, then there are ϕ(d) elements of order d. Proof. Let g have order n so that ...<|control11|><|separator|>
  27. [27]
    [PDF] Early group theory in the works of Lagrange, Cauchy, and Cayley
    Aug 19, 2010 · equation and α denote a primitive cubic root of unity. According to Lagrange's analysis in the preceding excerpt, the resolvent for the ...
  28. [28]
    Teoria generale delle equazioni : in cui si dimostra impossibile la ...
    May 19, 2009 · Teoria generale delle equazioni : in cui si dimostra impossibile la soluzione algebraica delle equazioni generali di grad superiore al quarto.
  29. [29]
    Paolo Ruffini's Contributions to the Quintic - jstor
    inclusion is proper, since otherwise we should have £l = £i_1 by a theorem of. Lagrange proved in Ruffini's "Teoria". Let ρ be a primitive 5th root of unity.
  30. [30]
    [PDF] ÉVARISTE GALOIS - Œuvres mathématiques - Numdam
    Dans la théorie des équations, j'ai recherché dans quels cas les équations étaient résolubles par des radicaux, ce qui m'a donné oc- casion d'approfondir cette ...
  31. [31]
    Évariste Galois: Principles and Applications
    ### Summary of Galois's 1831 Memoir and Primitive Element Theorem
  32. [32]
    [PDF] Galois Theory
    Jan 2, 2024 · ... element, called a primitive element (see Definition 3.1.6). Theorem 3.5.6. (Primitive element theorem). Let K be a field and E = K(β,γ)/K a ...
  33. [33]
    [PDF] THE GALOIS CORRESPONDENCE 1. Introduction Let L/K be a field ...
    Primitive Elements. Galois theory provides a method to prove a number is a primitive element in a Galois extension. Theorem 6.1. When L/K is a finite Galois ...Missing: Lagrange | Show results with:Lagrange
  34. [34]
    [PDF] arXiv:2408.02158v1 [math.NT] 4 Aug 2024
    Aug 4, 2024 · For part (i), the primitive element theorem implies that there exists an element ... to Hilbert's 12th Problem for such rational function fields.
  35. [35]
    [PDF] Class Field Theory
    He proved the following “unit theorem”: let ˛ be a root of a monic irreducible polynomial f . ... primitive element ˛ for L=K with ˛ 2 OL. Then. L D KŒ˛ ' KŒX =.
  36. [36]
    Introduction (NUMBER FIELDS) - Documentation
    An arbitrary number field can be converted to an absolute extension of Q using AbsoluteField, i.e. this finds a primitive element over Q. Similarly, a number ...
  37. [37]
    Computing primitive elements of extension fields - ScienceDirect.com
    Several mathematical results and new computational methods are presented for primitive elements and their minimal polynomials of algebraic extension fields.
  38. [38]
    [PDF] an effective version of the primitive element theorem
    In any Field Theory and Galois theory book, we first encounter the primitive element theorem. (see for instance, Theorem 4.6 in [6]) which states that if L/K is ...
  39. [39]
    Existence of pair of primitive elements over finite fields of ...
    Primitive elements find a lot of applications in cryptography and coding theory as they are generators of groups of non zero elements of finite fields.