Fact-checked by Grok 2 weeks ago

Rational point

In number theory and algebraic geometry, a rational point on an algebraic variety defined over the rational numbers \mathbb{Q} is a point whose coordinates all belong to \mathbb{Q}. More generally, for any field K, a K-rational point consists of coordinates in K, but the term "rational point" typically refers to the case where K = \mathbb{Q}, representing solutions to polynomial equations with rational values. The study of rational points seeks to determine their existence, distribution, and finiteness on varieties such as curves, surfaces, and higher-dimensional objects. For curves of 0, such as conics like the unit x^2 + y^2 = [1](/page/1), rational points are abundant and can be parameterized, corresponding to primitive Pythagorean triples like (3,4) or (5,12). Elliptic curves ( 1) form an under a geometric , enabling the infinite generation of rational points from a single one, as seen on y^2 = x^3 - 4x + [1](/page/1) with points like (0,[1](/page/1)) and (4,7). In contrast, Faltings' theorem (1983) proves that curves of at least 2 have only finitely many rational points, resolving long-standing conjectures like Mordell's for specific cases. Rational points hold profound significance beyond , underpinning applications in —such as for secure data transmission—and playing a pivotal role in proofs of major theorems, including ' resolution of via modular forms and elliptic curves. Counting rational points often involves height functions to bound their "size," with asymptotic growth analyzed as bounds increase, revealing density patterns over finite fields as well. Challenges persist in higher dimensions, where the tests local solvability over \mathbb{Q}_p and \mathbb{R} against global rational solutions, frequently failing due to obstructions like the Brauer-Manin.

Fundamentals

Definition

A k-rational point on an algebraic variety X defined over a field k is formally defined as a morphism \operatorname{Spec}(k) \to X in the category of schemes. Equivalently, for an affine variety X \subset \mathbb{A}^n_k defined by polynomial equations f_1 = \cdots = f_m = 0 with coefficients in k, a k-rational point corresponds to a solution (a_1, \dots, a_n) \in k^n satisfying these equations. This notion extends naturally to projective varieties using : a point on X \subset \mathbb{P}^n_k is represented by [x_0 : \dots : x_n] \in \mathbb{P}^n(k) with x_i \in k (not all zero) such that the homogeneous polynomials defining X vanish at these coordinates. Projective rational points are often preferred over affine ones because projective varieties are proper (compact in the classical topology when k = \mathbb{C}), ensuring better behavior under limits and avoiding issues with points "at infinity." In the more general setting of schemes, rational points are the k-points in the functor of points, i.e., X(k) = \operatorname{Hom}_k(\operatorname{Spec} k, X), where X is a separated k-scheme of finite type. In arithmetic geometry, "rational points" typically refer to absolute rational points over \mathbb{Q}, i.e., \mathbb{Q}-rational points on varieties defined over \mathbb{Q}. A basic example illustrates the distinction: the \mathbb{Q}-rational points on the affine line \mathbb{A}^1_\mathbb{Q} are simply \mathbb{Q}, consisting of points $$ for a \in \mathbb{Q}; however, on the projective line \mathbb{P}^1_\mathbb{Q}, they include the additional point at infinity [1:0].

Motivations and Historical Context

The study of rational points originates from ancient efforts to solve Diophantine equations, which seek integer or rational solutions to polynomial equations with integer coefficients. A classic early example is the search for Pythagorean triples—positive integers a, b, c satisfying a^2 + b^2 = c^2—corresponding to rational points on the unit circle x^2 + y^2 = 1. These triples, such as (3,4,5), were known to the Babylonians around 1800 BCE and later systematized by Euclid, illustrating how rational points parametrize solutions via stereographic projection from the point (-1,0). Central to number theory, rational points provide arithmetic invariants like class numbers and reveal the solvability of equations over the rationals \mathbb{Q}. Fermat's Last Theorem, conjectured in 1637, asserts no non-trivial positive integer solutions to x^n + y^n = z^n for n \geq 3, equivalent to the absence of non-trivial rational points on the Fermat curves x^n + y^n = z^n. This was proved by Andrew Wiles in 1995 using the modularity of semistable elliptic curves over \mathbb{Q}. Motivations deepened with Diophantine approximation, where Axel Thue's 1909 theorem bounded how well algebraic irrationals can be approximated by rationals, limiting solutions to equations like F(x,y) = 1 for binary forms F of degree at least 3. Carl Ludwig Siegel extended this in 1929, proving finiteness of integral points on curves of genus greater than zero using approximation techniques. Historical development accelerated with Louis Mordell's 1922 conjecture that rational points on elliptic curves (genus 1) form a , proved for \mathbb{Q} via descent methods. This Mordell-Weil theorem spurred broader inquiries into higher-genus curves. resolved the generalized Mordell in 1983, showing curves of genus at least 2 over number fields have finitely many rational points, employing the geometry of abelian varieties and heights. In the , computations by and on elliptic curves led to their linking the rank of the rational points group to the order of vanishing of the at s=1, connecting analytic and algebraic invariants. Modern relevance spans applications: the Mordell-Weil structure underpins elliptic curve cryptography, where group laws on curves over finite fields enable secure protocols like key exchange, building on rational point arithmetic.

Rational Points on Curves

Genus 0 Curves

Genus 0 curves, also known as rational curves, are projective curves of arithmetic genus 0. Over an algebraically closed field, such curves are isomorphic to the projective line \mathbb{P}^1. Over the rational numbers \mathbb{Q}, smooth genus 0 curves are precisely the conics, defined by homogeneous quadratic equations in three variables, such as x^2 + y^2 = z^2. A conic over \mathbb{Q} either has no rational points or infinitely many. If it has at least one rational point, then all rational points can be parametrized rationally, establishing a birational equivalence to \mathbb{P}^1 over \mathbb{Q}. The parametrization arises by considering lines through the known rational point with rational slope t; each such line intersects the conic at a second rational point. In projective coordinates (x:y:z), the rational points take the form (x:y:z) = (a t^2 + b t + c : d t^2 + e t + f : g t^2 + h t + i), where the coefficients depend on the fixed point and the conic equation. For the unit circle x^2 + y^2 = z^2, using the point (-1:0:1) and yields the parametrization (x:y:z) = (1 - t^2 : 2 t : 1 + t^2) for rational t, generating all primitive Pythagorean triples when clearing denominators. The existence of rational points on conics is governed by the Hasse principle, which holds in this case due to the Hasse-Minkowski theorem: a conic has a rational point if and only if it has points over the reals \mathbb{R} and over every p-adic field \mathbb{Q}_p. This local-global principle for quadratic forms in three variables was established by Hasse and Minkowski, resolving the solubility of equations like a x^2 + b y^2 + c z^2 = 0. A classical application is Legendre's three-square theorem, which states that a positive integer n is a sum of three integer squares if and only if it is not of the form $4^k (8m + 7); this criterion aligns precisely with the local solubility conditions over \mathbb{R} and all \mathbb{Q}_p. For example, the conic x^2 + y^2 = z^2 has infinitely many rational points, corresponding to Pythagorean triples like (3,4,5). In contrast, the conic x^2 + y^2 = 3 z^2 has no non-trivial rational points, as it fails the local condition modulo 3 (where -1 is not a ).

Genus 1 Curves

Elliptic curves over the rational numbers \mathbb{Q} are typically presented in Weierstrass form, given by the equation y^2 = x^3 + a x + b, where a, b \in \mathbb{Z} and the \Delta = -16(4a^3 + 27b^2) \neq 0 ensures the curve is nonsingular. The set of rational points E(\mathbb{Q}) on such a , including the point at \mathcal{O}, forms an under the chord-and-tangent addition law: the sum of two points P and Q is the reflection across the x-axis of the third intersection point of the line through P and Q with the curve, and multiples are defined iteratively. The Mordell-Weil theorem asserts that E(\mathbb{Q}) is a finitely generated abelian group, isomorphic to \mathbb{Z}^r \oplus T, where r \geq 0 is the rank and T is the finite torsion subgroup. This result, originally proved by Mordell for curves of this form in 1922, was generalized by Weil to elliptic curves over any number field. The torsion subgroup T is finite, and Mazur's theorem classifies all possible structures: T is either \mathbb{Z}/N\mathbb{Z} for N = 1 to $10 or $12, or \mathbb{Z}/2\mathbb{Z} \oplus \mathbb{Z}/2N\mathbb{Z} for N = 1 to $4. Representative examples illustrate these structures. The curve y^2 = x^3 - 2 has and trivial torsion, generated by the point (3, 5). In contrast, the curve y^2 + y = x^3 - x has and trivial torsion, with generator (0, 0). While the Hasse principle holds for genus 0 curves, it fails for some elliptic curves: Selmer constructed in the curve $3x^3 + 4y^3 + 5z^3 = 0, which has points over \mathbb{R} and every \mathbb{Q}_p but no rational points. Such failures are quantified by the Tate-Shafarevich group \Sha(E/\mathbb{Q}), the kernel of the map from the group of principal homogeneous spaces to the product over places of local solutions; it measures global obstructions beyond local solvability. The Birch-Swinnerton-Dyer conjecture predicts that |\Sha(E/\mathbb{Q})| is finite and equals the leading Taylor coefficient of the L(E/\mathbb{Q}, s) at s=1 divided by the regulator times the Tamagawa factors and torsion order. Recent advances include partial finiteness results for \Sha in specific cases, such as over function fields for algebraic tori using adelic methods, and growth bounds for p-primary components in supersingular abelian varieties via . Euler systems have yielded explicit bounds on \Sha for elliptic curves of analytic rank at most 1, with post-2020 refinements addressing visibility in modular quotients.

Higher Genus Curves

For curves of genus g \geq 2 defined over a number field K, Faltings's theorem establishes the finiteness of the set of rational points C(K). Specifically, if C is a smooth projective curve of genus at least 2 over K, then C(K) is finite. This result, proved in 1983, resolves the Mordell conjecture in the affirmative for higher genus. The proof relies on the geometry of the moduli space of stable curves and abelian varieties, combined with bounds on Néron-Tate heights to control the distribution of points and show that only finitely many can exist without violating height inequalities. Hyperelliptic curves provide concrete examples of this finiteness. A hyperelliptic curve of genus 2 over \mathbb{Q} takes the form y^2 = f(x), where f(x) is a square-free polynomial of degree 5 or 6; Faltings's theorem implies that such a curve has only finitely many rational points. For instance, the curve y^2 = x^5 + 1 has genus 2 and exactly three rational points: (0, \pm 1) and the point at infinity. Modular curves of higher genus, such as X_0(21) which has genus 2, also illustrate sparse rational points; computations show it possesses precisely six rational points, corresponding to cusps and specific elliptic curves with complex multiplication. These examples highlight how the theorem applies to families with explicit geometric structure. Effective versions of provide quantitative bounds on the number of rational points. For 2 curves over \mathbb{Q}, methods combining the Chabauty-Coleman technique with Mordell-Weil sieving yield explicit upper bounds, often determining all points computationally for curves with small conductors. More ambitiously, Vojta's conjectures predict bounds depending only on the and the degree of the number field, independent of the specific ; partial progress toward this uniform Mordell conjecture includes bounds in terms of the , a degree parameter, and the Mordell-Weil rank of the , as established in recent work. These refinements enable practical computations and strengthen connections to broader diophantine problems. The Mordell conjecture, posed in 1922 for curves over \mathbb{Q}, posited finiteness of rational points for genus greater than 1 and was generalized to number fields; Faltings's proof marked a landmark in arithmetic geometry. The theorem has applications to the through constructions involving Frey curves, where finiteness constraints on higher-genus twists inform bounds on radical differences in Diophantine equations.

Rational Points on Higher-Dimensional Varieties

Varieties with Sparse Rational Points

In varieties of general type, rational points are expected to be sparse, meaning they form a set that is not Zariski dense and often lies on proper subvarieties. The Bombieri–Lang conjecture, proposed in the 1980s, asserts that for a smooth projective variety of general type defined over a number field, the rational points are contained in a proper Zariski closed subset, with the exceptional locus itself of general type. This prediction generalizes Faltings's theorem for curves of genus at least 2, where rational points are finite, to higher dimensions. Examples include high-degree hypersurfaces in projective space, such as quintic hypersurfaces in \mathbb{P}^4, where the conjecture implies that rational points cannot be dense but must reside on lower-dimensional subvarieties of lower Kodaira dimension. An addendum by specifically addresses hypersurfaces, conjecturing that for hypersurfaces of sufficiently high degree in \mathbb{P}^n (with n \geq 3), the rational points over a number field lie on a proper subvariety, reinforcing the sparsity for varieties of general type. Faltings extended his methods beyond curves to prove finiteness results for rational points on certain higher-dimensional subvarieties. In particular, for a subvariety of an that is not a translate of an abelian subvariety, the rational points are finite if the subvariety has (i.e., is of general type). This applies, for instance, to general-type subvarieties embedded in abelian varieties, yielding only finitely many rational points outside torsion translates. Specific cases illustrate these sparsity phenomena. For cubic hypersurfaces in \mathbb{P}^4, which are Fano threefolds of index 1, Manin's conjecture predicts the asymptotic distribution of rational points of bounded height as c B^{2} (\log B)^{r-1} (where r relates to the rank of the intermediate Jacobian), indicating controlled growth rather than density; partial proofs confirming leading terms have appeared since 2010 for singular cases and refined estimates for smooth ones. On K3 surfaces, which have Kodaira dimension 0 but nontrivial canonical bundle, rational points are conjectured not to be Zariski dense, often confined to curves or finite sets under conditions like high Picard rank, aligning with broader Vojta-type expectations for non-general-type cases. The Caporaso–Harris–Mazur theorem from 1997 establishes that the Bombieri–Lang conjecture implies uniform boundedness on the number of rational points for curves of fixed g \geq 2 over number fields of bounded degree, providing a quantitative sparsity measure. Recent progress toward uniform finiteness in higher dimensions draws from Vojta's program, which posits height inequalities implying Bombieri–Lang; post-2020 developments include formulations and partial verifications of Vojta's conjectures for weighted projective varieties and log pairs, supporting uniform bounds on rational points for families of general-type varieties.

Varieties with Dense Rational Points

In higher-dimensional varieties over the rationals, rational points can form infinite sets that are dense in the , particularly for those exhibiting certain geometric structures such as toric varieties or fibrations with rational fibers. For smooth projective split toric varieties over \mathbb{Q}, the set of rational points is Zariski dense whenever it is nonempty, as these varieties are quasi-Fano and admit torsors facilitating the distribution of points via functions. Similarly, varieties admitting fibrations over \mathbb{Q} with fibers that are rational or have dense rational points inherit this abundance, allowing rational sections to generate dense subsets through fiberwise parametrization. A unifying framework for such phenomena is provided by the notion of special varieties, introduced by Campana in the context of orbifold classification theory. A compact is special if it admits no dominant meromorphic map onto an of general type; over number fields like \mathbb{Q}, special varieties are conjectured to possess potentially dense rational points, meaning dense after a finite extension. This conjecture, formulated in the and known in low dimensions (up to 4) and specific higher-dimensional cases such as certain fibrations, posits that geometric "specialness"—absence of general type components—implies arithmetic abundance of rational points. Projective space \mathbb{P}^n over \mathbb{Q} exemplifies this density: its rational points, consisting of homogeneous coordinates in \mathbb{Q}^{n+1} \setminus \{0\} modulo scaling, form a Zariski dense subset, as they intersect every nonempty open set and their height-bounded counts grow asymptotically like B^{n+1}. Abelian varieties over \mathbb{Q} generalize the Mordell-Weil theorem from elliptic curves (dimension 1, where infinite rational points are Zariski dense on the curve) to higher dimensions, yielding a finitely generated group of rational points that is infinite when the rank is positive, though Zariski density requires the rank to span the dimension sufficiently. Cubic surfaces over \mathbb{Q}, which are del Pezzo surfaces of degree 3, provide another concrete example: a smooth possessing at least one rational point is birational to \mathbb{P}^2 over \mathbb{Q}, ensuring its rational points are Zariski dense, with counts of bounded height growing like R^{3/2 + \epsilon} under rank hypotheses on associated elliptic curves. More broadly, for unirational varieties over \mathbb{Q}—those admitting a dominant rational map from projective space—the existence of one rational point implies the rational points are Zariski dense, as the image of \mathbb{Q}-points under the parametrization fills the variety densely. Surfaces illustrate density results tied to curvature bounds: the Bogomolov-Miyaoka-Yau , stating c_1^2 \leq 3c_2 for minimal surfaces of general type, demarcates those with limited rational points from non-general type surfaces (e.g., rational or ruled), which admit ample rational curves and thus dense rational points when defined over \mathbb{Q}. For del Pezzo surfaces, recent work confirms potential of rational points over non-closed fields, with explicit moduli constructions showing Zariski after finite extensions for degrees up to 4.

Arithmetic Tools and Methods

Point Counting over Finite Fields

In algebraic geometry, counting points on varieties over finite fields serves as a powerful analogy for understanding the distribution of rational points over the rationals, particularly through heuristics involving local densities at primes. For a variety X defined over the finite field \mathbb{F}_q, the cardinality \#X(\mathbb{F}_q) denotes the number of \mathbb{F}_q-rational points, which are tuples of coordinates in \mathbb{F}_q satisfying the defining equations of X. These counts capture the "local" behavior of X modulo primes and inform predictions about the global density of rational points via the product of local densities, providing insights into whether varieties admit infinitely many or finitely many rational points. The , formulated by in the late 1940s, establish a rigorous framework for these point counts, drawing parallels to the for number fields. For a smooth X of dimension d over \mathbb{F}_q, the conjectures posit that the zeta function Z(X, t) = \exp\left( \sum_{k=1}^\infty \#X(\mathbb{F}_{q^k}) \frac{t^k}{k} \right) (with t = q^{-s}) is a satisfying a , with poles and zeros at specified locations determined by the ; moreover, the asserts that the reciprocals of the roots (eigenvalues of geometric Frobenius) have absolute value q^{w/2} in the w-th group, leading to an error term bounded by the Betti numbers in the point count formula \#X(\mathbb{F}_q) = \sum_{i=0}^{2d} (-1)^i \operatorname{Tr}(F_q^* | H^i_c(\overline{X}, \mathbb{Q}_\ell)). For curves of g, this simplifies to the explicit Hasse-Weil bound |\#X(\mathbb{F}_q) - (q + 1)| \leq 2g \sqrt{q}, which quantifies the deviation from the "expected" number of points q + 1. The bound for elliptic curves (g=1) was proved by in 1933 using class number estimates for imaginary quadratic fields, while the full , including the general curve case, were proved by in 1974 using and the hard Lefschetz theorem. These finite field counts have significant applications in arithmetic geometry, particularly through Hasse-Weil L-functions that interpolate the data to connect local information to global rational point structures. For an elliptic curve E over \mathbb{Q}, the Hasse-Weil L-function is the Euler product L(E, s) = \prod_p L_p(E, s)^{-1}, where the local factor at an unramified prime p is L_p(E, s) = 1 - a_p p^{-s} + p^{1-2s} with trace a_p = p + 1 - \#E(\mathbb{F}_p); this L-function admits analytic continuation to \mathbb{C} and satisfies a functional equation, as predicted by the Weil conjectures and verified via modularity. The Birch and Swinnerton-Dyer conjecture links this to rational points by asserting that the order of vanishing of L(E, s) at s=1 equals the rank of the Mordell-Weil group E(\mathbb{Q}), with the leading term involving the regulator, Tamagawa numbers, and the order of the torsion subgroup; partial evidence, such as for ranks 0 and 1, stems from the correspondence between point counts and modular forms. Additionally, the Katz-Lang method employs uniform bounds on point counts over finite fields—derived from the Weil conjectures—to prove finiteness theorems for rational points in unlikely intersections, such as showing that the intersection of a curve with a proper algebraic subgroup of an abelian variety contains only finitely many rational points under suitable height conditions. A representative example arises for elliptic curves, where the point count formula \#E(\mathbb{F}_p) = p + 1 - a_p holds for odd primes p of good reduction, with the Hasse bound |a_p| \leq 2\sqrt{p} ensuring the traces remain controlled and enabling the construction of the L-function; this bound has been pivotal in computational verification of ranks and in cryptographic applications, illustrating the quantitative insight from finite fields into rational point scarcity or abundance.

Height Functions and Effective Bounds

Height functions provide a quantitative measure of the arithmetic of on algebraic varieties, facilitating effective bounds in . For a point P = [x_0 : \cdots : x_n] \in \mathbb{P}^n(\mathbb{Q}) in , the (absolute logarithmic) Weil height is defined as h(P) = \log \max_i |x_i|, where the coordinates are integers in lowest terms with \gcd(x_0, \dots, x_n) = 1. This extends to points over number fields K via the product formula over places, yielding h_K(P) = \frac{1}{[K:\mathbb{Q}]} \sum_{v \in M_K} \log \max_i \|x_i\|_v, which captures the "size" of the point in a geometrically way. On varieties such as or , canonical heights refine this notion to align with the group structure. For an E over a number field K, the Néron-Tate canonical \hat{h} on the Mordell-Weil group E(K) satisfies \hat{h}(P) \geq 0 with equality if and only if P is torsion, and exhibits quadratic behavior under the group law: \hat{h}(nP) = n^2 \hat{h}(P) for integer multiplication-by-n. It is constructed as the \hat{h}(P) = \lim_{n \to \infty} \frac{1}{n^2} h(P), differing from the Weil by a bounded amount. A key property is Northcott's theorem, which asserts that for fixed and , there are only finitely many points in \mathbb{P}^n(K) (or on a ) of bounded , implying finiteness of rational points of controlled complexity. These heights enable effective bounds on Diophantine problems. posits that for an E/\mathbb{Q} with minimal discriminant \Delta and conductor N, |\Delta| \ll_\epsilon N^{6+\epsilon} for any \epsilon > 0, linking conductor growth to height-like measures and equivalent to the weak via Frey curves. For Mordell-Weil ranks, Silverman's bounds yield r \leq C_g d^3 \max\{1, h_F^+(E/K), \log |\Delta_K|\}, where r is the rank, g=1 for elliptic curves, d = [K:\mathbb{Q}], h_F^+ is the Faltings height (related to canonical heights), and C_g is an absolute constant, providing explicit control via arithmetic invariants. Applications include effective versions of finiteness theorems. An effective Shafarevich theorem bounds the heights of minimal Weierstrass models for elliptic curves over K with good reduction outside a S of places: the coefficients satisfy \max(h(a_4), h(a_6)) \leq \exp(\exp(C (s + h_K \log |D_K| + \log \log p)^2)), where s = |S|, h_K is the class number, D_K the , and p the largest prime in S, making the finiteness computable. Post-2020 advances have strengthened uniform bounds on rational points of bounded height; for instance, on elliptic curves over number fields with a rational point of prime \ell, the number of such points is at most B^{C / \log \log B} for canonical height \leq \log B, with C effective depending on K and \ell, leveraging relative for sharper estimates. These tools, often informed by o-minimal structures for in families, address longstanding uniformity questions in rational point distribution.

Obstructions to Rational Points

Local-Global Principles

The Hasse principle, also known as the local-global principle, asserts that a X defined over the rational numbers \mathbb{Q} has a rational point if and only if it has points over numbers \mathbb{R} and over the p-adic fields \mathbb{Q}_p for every prime p. This principle holds for varieties of genus 0, such as conics, where local solvability guarantees a global rational point. However, it fails in general for higher-degree varieties, as demonstrated by various counterexamples. Local solvability is formalized in terms of adelic points: the adele ring \mathbb{A}_\mathbb{Q} is the restricted product of the local fields \mathbb{Q}_v over all places v of \mathbb{Q}, and X(\mathbb{A}_\mathbb{Q}) = \prod_v' X(\mathbb{Q}_v) consists of tuples of local points compatible away from finitely many places. The Hasse principle states that X(\mathbb{Q}) \neq \emptyset if and only if X(\mathbb{A}_\mathbb{Q}) \neq \emptyset. A stronger property, weak approximation, holds when the image of X(\mathbb{Q}) under the diagonal embedding into X(\mathbb{A}_\mathbb{Q}) is dense in the , which occurs for varieties like but fails for elliptic curves in general. A classic counterexample is Selmer's curve defined by the equation $3x^3 + 4y^3 + 5z^3 = 0 in over \mathbb{[Q](/page/Q)}, which has points over \mathbb{R} and every \mathbb{Q}_p but no rational points. This curve is of genus 1 and provides an early illustration of the principle's failure for cubic equations. For genus 1 curves more broadly, failures of the Hasse principle are quantified by the Tate-Shafarevich group, as explored in Cassels' foundational work on the arithmetic of such curves. Variants of the principle succeed in specific contexts, such as strong approximation for the \mathrm{SL}_n over \mathbb{Q}, where the image of \mathrm{SL}_n(\mathbb{Z}) is dense in \mathrm{SL}_n(\mathbb{A}_\mathbb{Q}^S) for any S of places, extending the Hasse principle to simultaneous control over local conditions. Similarly, the Grunwald-Wang theorem addresses the local-global principle for cyclic algebras over number fields, guaranteeing the existence of a cyclic algebra with prescribed local invariants at a of places, except in a special case involving 8th roots of unity.

Brauer-Manin Obstruction

The Brauer-Manin obstruction provides a cohomological tool to detect failures of the Hasse principle for the existence of rational points on varieties over number fields. For a smooth projective variety X defined over a number field k, the Brauer group \mathrm{Br}(X) is defined as the torsion subgroup of the étale cohomology group H^2_{\ét}(X_{\overline{k}}, \mathbb{G}_m), which is isomorphic to H^2_{\ét}(X, \mathbb{Q}/\mathbb{Z}) under suitable conditions. The obstruction arises from a pairing \mathrm{Br}(X) \times X(\mathbb{A}_k) \to \mathbb{Q}/\mathbb{Z}, where \mathbb{A}_k denotes the adele ring of k. This pairing is given by \langle A, P \rangle = \sum_v \mathrm{inv}_v(A)(P_v) for A \in \mathrm{Br}(X) and P = (P_v)_v \in X(\mathbb{A}_k), with \mathrm{inv}_v the local invariant map at each place v of k. The Brauer-Manin set is then X(\mathbb{A}_k)^{\mathrm{Br}} = \{ P \in X(\mathbb{A}_k) \mid \langle A, P \rangle = 0 \ \forall A \in \mathrm{Br}(X) \}. If this set is empty, then X(k) = \emptyset, as any rational point would embed diagonally into X(\mathbb{A}_k)^{\mathrm{Br}}. This obstruction refines local-global principles by explaining certain adelic points without global counterparts through non-trivial Brauer classes. The concept was introduced by in his 1970 International Congress of Mathematicians address, where he applied it to conic bundles and pencils of conics over number fields, showing that non-trivial elements in the Brauer group obstruct rational points despite local solvability everywhere. Later applications extended this to higher-degree varieties; for instance, computations on del Pezzo surfaces by Alexander Skorobogatov demonstrated that the Brauer-Manin obstruction accounts for many Hasse principle violations, particularly on degree-4 del Pezzo surfaces where transcendental Brauer classes play a key role. In the context of genus-1 curves, the obstruction explains failures in Selmer's classical list of 16 examples violating the Hasse principle, linking them to non-trivial 2-torsion in the Brauer group derived from the Jacobian's Tate-Shafarevich group. Despite its explanatory power, the Brauer-Manin obstruction is not always sufficient to account for all Hasse principle failures. Skorobogatov's 1999 construction of bielliptic surfaces, and later examples involving quadratic twists of genus-2 curves, reveal varieties locally solvable everywhere but with nonempty Brauer-Manin set yet still lacking rational points, indicating additional obstructions beyond the Brauer group. Recent computational advances, such as the 2023 work by Stoll et al. on conic bundles, have shown that the obstruction can require arbitrarily many generators to capture the full effect in some instances by computing the full Brauer group. The Brauer-Manin obstruction connects to cohomological via the Tate-Shafarevich group \Sha(E) of the E of a , where compatible systems of torsors under finite étale covers correspond to elements in \Sha, and the obstruction measures whether these systems admit global data obstructed by Brauer classes. This link, developed in works by Colliot-Thélène and Sansuc, shows that for principally polarized abelian varieties, the obstruction often captures the full obstruction when \Sha is finite.

References

  1. [1]
    Rational Point -- from Wolfram MathWorld
    A -rational point is a point on an algebraic curve , where and are in a field . For example, rational point in the field of ordinary rational numbers is a ...
  2. [2]
    [PDF] Introduction to rational points
    Jan 17, 2006 · number of points in X(Q) whose coordinates have numerator and denominator bounded by B in absolute value, and see how this count grows as B → ∞.
  3. [3]
    How to Find Rational Points Like Your Job Depends on It
    Jul 22, 2021 · “Rational points” are points in the plane whose coordinates are all rational numbers. For example, \left(\frac{12}{5},-\frac{2}{3}\right), \left ...<|control11|><|separator|>
  4. [4]
    [PDF] Rational points on varieties - MIT Mathematics
    Jun 6, 2010 · ... k into varieties over k. Chapter 5 is a survey on group schemes and algebraic groups over fields. After discussing their general properties ...
  5. [5]
    [PDF] pythagorean triples - keith conrad
    A Pythagorean triple is a triple of positive integers (a, b, c) where a² + b² = c². Examples include (3,4,5), (5,12,13), and (8,15,17).
  6. [6]
    [PDF] Modular elliptic curves and Fermat's Last Theorem
    The object of this paper is to prove that all semistable elliptic curves over the set of rational numbers are modular. Fermat's Last. Theorem follows as a ...
  7. [7]
    [PDF] Birch-Conjectures_Concerning_Elliptic_Curves.pdf
    I want to describe some computations undertaken by myself and Swinnerton-Dyer on EDSAC, by which we have calculated the zeta-functions of certain elliptic ...
  8. [8]
    [hep-th/9202017] Rational Curves on Calabi-Yau Threefolds - arXiv
    Feb 5, 1992 · The point of this talk is to give mathematical techniques and examples for computing the finite number that ``should'' correspond to an infinite ...
  9. [9]
    [PDF] Rational points on curves - Math (Princeton)
    Dec 6, 2013 · 1.1 Rational points. Let C be a curve of genus 0 defined over rational. We are concerning the question when C has a rational point in Q.
  10. [10]
    [PDF] Rational Points on Conics, Lecture 24 Notes - MIT OpenCourseWare
    Principle - if we can find one (rational) point on a sonic, then we can parametrize all rational points, and there are infinitely many of them.
  11. [11]
    [PDF] rational parametrization of conics
    A K-rational point of L is a solution. (x, y) ∈ K2 of L. The set of K-rational points of L is denoted LK. Similarly, a conic curve defined over k is an equation.
  12. [12]
    [PDF] The Hasse-Minkowski Theorem - Digital Commons @ UConn
    Mar 8, 2006 · Thus the task of finding all the rational points on a diagonal plane conic over Q boils down to knowing the existence of one rational point.
  13. [13]
    [PDF] Rational Points on Conics, and Local-Global Relations in Number ...
    Nov 26, 2007 · Legendre's theorem has evolved into the Hasse principle, relating. “local” and “global” solutions, even for more variables. This will all be.
  14. [14]
    [PDF] Legendre's theorem, LEGRANGE'S DESCENT - CSUSM
    Corollary 7 (Hasse Principle: form 3). Consider a quadratic homogeneous Dio- phantine equation F(X,Y,Z) = 0 where F(X,Y,Z) ∈ Z[X,Y,Z] has degree 2. Then ...
  15. [15]
    [PDF] PROOF THAT x2 + y2 = 3 HAS NO RATIONAL SOLUTIONS.
    Here, we will use a similar framework to explain why x2 +y2 = 3 has no solutions with x, y ∈ Q. In what follows, we will try to highlight the main structural ...
  16. [16]
    [PDF] Rational isogenies of prime degree - Columbia Math Department
    An immediate application of the above corollary is the following theorem which classifies all possible torsion groups of Mordell-Weil groups of elliptic curves ...
  17. [17]
    [PDF] The finiteness of the Tate–Shafarevich group over function fields for ...
    In Section 2, we provide a short summary of our adelic approach to the finiteness of Tate–Shafarevich groups of tori over finitely generated fields with respect ...
  18. [18]
    [PDF] Rational points on curves
    Apr 11, 2008 · Theorem 2.1 (Faltings). Let X be a smooth projective curve of genus ≥ 2 defined over a number field K. Then X(K) is finite.<|control11|><|separator|>
  19. [19]
    [PDF] Rational points on curves
    Sep 18, 2015 · Theorem 1.1 (Faltings [Fal83]). If C is a smooth, projective and absolutely irreducible curve over Q of genus g ≥ 2, then C(Q) is finite.
  20. [20]
    [PDF] A database of genus 2 curves over the rational numbers - arXiv
    We find a total of 66,158 isomorphism classes of curves with absolute discriminant at most 106; for each curve, we compute an array of geometric and arithmetic.
  21. [21]
    Rational points on the modular curves $X_{0}^{+}(N) - Project Euclid
    is isomorphic over $C$ to $(E^{\sigma}, A^{\sigma})$ . There are rational points on $X_{0}^{+}(N)$ which are represented by elliptic curves.
  22. [22]
    [PDF] Determining the rational points on a curve of genus 2 and Mordell ...
    Sep 29, 2025 · Use a rational divisor of odd degree to embed C into J and run a “Mordell-Weil sieve. + Chabauty” computation. This will terminate in practice ( ...
  23. [23]
    [PDF] Computational aspects of curves of genus at least 2 - MIT Mathematics
    (He needed this for his work with Baker on effective bounds for integer points on elliptic curves [7].) Much more recently, Huang and Ierardi [50] proved that ...
  24. [24]
    Uniformity in Mordell–Lang for curves - Annals of Mathematics
    We show that the number of rational points is bounded only in terms of g, d, and the Mordell–Weil rank of the curve's Jacobian.
  25. [25]
    [PDF] Modular forms, elliptic curves, and the ABC conjecture
    In 1991 Elkies [E] proved that the ABC–conjecture implies the Mordell conjecture (this was first proved by Faltings [F]) which states that every algebraic ...
  26. [26]
    The Erdos-Ulam problem, varieties of general type, and ... - Terry Tao
    Dec 20, 2014 · In fact, the Bombieri-Lang conjecture has been made for varieties of arbitrary dimension, and for more general number fields than the rationals ...
  27. [27]
    Lang's conjecture beyond the curve case - MathOverflow
    Dec 23, 2020 · If V is defined over a number field K, then one has the following conjecture due to Lang (Bombieri had made a similar conjecture in the case of ...Missing: addendum | Show results with:addendum
  28. [28]
    Diophantine approximation on abelian varieties
    Diophantine approximation on abelian varieties. Pages 549-576 from Volume 133 (1991), Issue 3 by Gerd Faltings. No abstract available for this article.
  29. [29]
    Manin's conjecture for a class of singular cubic hypersurfaces - arXiv
    Mar 17, 2017 · We establish an asymptotic formula for the number of rational points of bounded height on singular cubic hypersurfaces S_n defined by x^3=(y_1^2 + \cdots + y_n ...Missing: P^ | Show results with:P^
  30. [30]
    K3 surfaces with Picard number one and infinitely many rational points
    They asked whether there exists a K3 surface over a number field and with Picard number 1 that contains infinitely many rational points. In this article we will ...Missing: finiteness | Show results with:finiteness
  31. [31]
    Uniformity of Rational Points. - American Mathematical Society
    L.Caporaso, J.Harris, B.Mazur. How many rational points can a curve have? Proceed- ings of the Texel Conference, Progress in Math. vol. 129, Birkhauser ...Missing: boundedness | Show results with:boundedness
  32. [32]
    [2309.10300] Vojta's conjecture on weighted projective varieties
    Sep 19, 2023 · We formulate Vojta's conjecture for smooth weighted projective varieties, weighted multiplier ideal sheaves, and weighted log pairs.
  33. [33]
    [PDF] THE WEIL CONJECTURE. I
    Oct 24, 2021 · In this article, I prove the Weil conjecture on the eigenvalues of Frobenius endomor- phisms. The precise statement is given in (1.6).
  34. [34]
    [PDF] THE BIRCH AND SWINNERTON-DYER CONJECTURE
    One very old problem concerned with rational points on elliptic curves is the congruent number problem. One way of stating it is to ask which rational integers.
  35. [35]
    [PDF] An Introduction to Height Functions - MSRI
    Jan 21, 2006 · The canonical height allows us to accurately count the rational points of bounded height on abelian varieties. Theorem. (Néron) Let D be an ...
  36. [36]
    [PDF] Canonical Heights on Abelian Varieties Lecture Notes for the ...
    Further, we have defined traces and stated Theorem 3.25 when K = k(C) is the function field of a curve, but this material can be extended to higher dimensional ...
  37. [37]
    [1506.05165] Heights, ranks and regulators of abelian varieties - arXiv
    Jun 16, 2015 · It leads to an unconditional bound on the rank of Mordell-Weil groups. Assuming the height conjecture of Lang and Silverman, we then obtain ...
  38. [38]
  39. [39]
    [PDF] Strong approximation for algebraic groups - The Library at SLMath
    This article is a survey of known results related to strong approximation in algebraic groups. We focus primarily on two aspects: the classical form of strong.
  40. [40]
    [PDF] The Grunwald-Wang theorem
    May 1, 2013 · In section 6 we finally prove the Grunwald-Wang theorem by reducing it to the Hasse principle for mth powers via class field theory and.
  41. [41]
    [PDF] local-global principle for rational points and zero-cycles arizona ...
    The Hasse principle fails for a k-variety if Qv∈Ω. X(kv) 6= ∅ and X(k) = ∅. A class of algebraic varieties over k satisfies the Hasse principle if any k-variety.<|control11|><|separator|>
  42. [42]
    [PDF] Rational points on varieties and the Brauer-Manin obstruction
    They give my perspective on how the feedback loop of computation and theory currently manifests in the study of rational points and the Brauer-Manin obstruction ...
  43. [43]
    [PDF] Heuristics for the Brauer-Manin obstruction for curves
    Brauer-Manin obstruction. For the connection of the Brauer-Manin obstruction to the information on rational points obtained from finite étale covers, see ...
  44. [44]
    [PDF] Insufficiency of the Brauer-Manin obstruction applied to étale covers
    May 3, 2010 · But the insuf- ficiency was proved only in 1999, when a ground-breaking paper of Skoroboga- tov [Sko99] constructed a variety for which one ...
  45. [45]
    [PDF] Brauer-Manin obstructions requiring arbitrarily many Brauer classes
    Aug 15, 2023 · This project started at the Park City Mathematics Institute (PCMI) 2022 program “Num- ber theory informed by computation”. We thank the PCMI ...Missing: post- | Show results with:post-