Fact-checked by Grok 2 weeks ago

Plane curve

A plane curve is a curve that lies entirely within a two-dimensional and can be described either as the set of points (x(t), y(t)) traced by continuous x(t) and y(t) over an interval of the parameter t, or implicitly as the zero locus of a f(x, y) = 0. Plane curves form a foundational object in both and , enabling the study of geometric properties such as , , and in the parametric case, where the at a point is given by \frac{dy/dt}{dx/dt} and the \kappa measures the rate of turning via \kappa = \left| \frac{d\mathbf{T}/dt}{ds/dt} \right| with parameter s. For algebraic plane curves, defined by equations f(x, y) = 0 of d, key invariants include the d—with lines at degree 1 and conics at degree 2—and singularities, points where the curve fails to be , affecting properties like the . Notable examples of plane curves include straight lines, parametrized as x(t) = r \cos t, y(t) = r \sin t, and parabolas like y = x^2, alongside more complex forms such as cycloids generated by a rolling , x(t) = a(t - \sin t), y(t) = a(1 - \cos t), which illustrate applications in . In , states that two plane curves of degrees m and n intersect at exactly mn points in the , counting multiplicities, underpinning . The study of plane curves originated with conic sections and evolved through 19th-century works on higher-degree curves, such as George Salmon's treatise, influencing modern fields like and .

Representations

Parametric Representation

A parametric plane curve is defined as a mapping \mathbf{r}(t) = (x(t), y(t)) from an I \subseteq \mathbb{R} to \mathbb{R}^2, where t is the and x(t), y(t) are continuous functions. This traces the curve as t varies over I, providing a flexible way to describe paths in the plane. Parametric forms offer advantages over explicit or implicit representations, particularly for closed curves, self-intersecting curves, and multi-valued relations that cannot be easily expressed as single-valued functions of one . For instance, they symmetrically treat x and y, simplifying computations and extending naturally to complex shapes without restrictions on functionality. To convert a parametric curve to Cartesian coordinates, eliminate the parameter t by solving one equation for t and substituting into the other, yielding a F(x, y) = 0. A classic example is the unit circle, given by x(t) = \cos t, y(t) = \sin t for t \in [0, 2\pi), which eliminates to x^2 + y^2 = 1. The arc length L of a parametric curve from t = a to t = b is given by L = \int_a^b \sqrt{ \left( \frac{dx}{dt} \right)^2 + \left( \frac{dy}{dt} \right)^2 } \, dt, assuming the derivatives exist and the integrand is continuous. The orientation of the curve, or the direction of traversal, is determined by the direction in which t increases, such as counterclockwise for example as t rises from 0 to $2\pi.

Implicit Representation

An implicit plane curve is defined as the zero set of a f: \mathbb{R}^2 \to \mathbb{R}, consisting of all points (x, y) in the plane satisfying f(x, y) = 0. This representation treats the curve as a locus of points where the vanishes, without specifying a parameterization or solving for one variable in terms of the other. Such equations are particularly suited for algebraic manipulations and symmetric descriptions of curves, as they relate x and y directly. The gradient vector \nabla f = \left( \frac{\partial f}{\partial x}, \frac{\partial f}{\partial y} \right) at on the provides to the , assuming \nabla f \neq (0, 0). This allows implicit determination of the as to the . Common examples include the , defined by x^2 + y^2 - 1 = 0, where \nabla f = (2x, 2y), and , given by ax + by + c = 0, with (a, b) constant along the line. To find the slope of the , implicit yields \frac{dy}{dx} = -\frac{\frac{\partial f}{\partial x}}{\frac{\partial f}{\partial y}} at points where \frac{\partial f}{\partial y} \neq 0, treating y as a of x along the . For the unit circle, this gives \frac{dy}{dx} = -\frac{x}{y}. However, implicit equations can describe curves with multiple connected components, such as disconnected loops, or singular points where the vanishes, complicating . Parameterizing these curves for explicit traversal often requires solving numerically or algebraically.

Explicit Representation

An explicit plane curve is defined as the y = f(x), where f is a real-valued defined on some or in the real numbers, representing the set of points (x, f(x)) in the Cartesian plane. This form provides a direct vertical mapping from x-coordinates to y-coordinates, making it particularly suitable for visualizing and analyzing curves that are single-valued with respect to the x-axis. However, this representation has notable limitations. It cannot capture curves with vertical tangents, such as a vertical line x = c, because no y = f(x) can assign multiple y-values to a single x or handle undefined slopes at vertical points. Additionally, it fails to represent closed loops, like a , or multi-branched curves, where a single x might correspond to multiple y-values, requiring piecewise definitions or alternative approaches. A key advantage in calculus applications is that the derivative f'(x) directly provides the slope of the tangent line to the curve at any point (x, f(x)), facilitating computations of rates of change and linear approximations. For instance, the equation of the tangent line at x = a is y - f(a) = f'(a)(x - a). The arc length L of such a curve from x = a to x = b, assuming f is differentiable and f' is continuous on [a, b], is given by the integral L = \int_{a}^{b} \sqrt{1 + \left( \frac{dy}{dx} \right)^2} \, dx = \int_{a}^{b} \sqrt{1 + [f'(x)]^2} \, dx. This formula arises from approximating the curve with line segments and taking the limit, providing a precise measure of the curve's total length along its path. A classic example is the parabola y = x^2, defined for all real x, which traces a U-shaped curve symmetric about the y-axis. This explicit form can be converted to an implicit equation by rearranging to x^2 - y = 0, highlighting its algebraic nature while preserving the curve's geometry.

Smooth Plane Curves

Definition and Differentiability

In , a plane curve is typically studied through its parametric representation as a map \mathbf{r}: I \to \mathbb{R}^2, where I \subset \mathbb{R} is an open interval, allowing the application of differentiability concepts to analyze its geometric properties. A smooth plane curve is defined as a C^\infty map from I to \mathbb{R}^2, meaning all derivatives exist and are continuous, which ensures the curve has no cusps or corners in its local structure. This C^\infty smoothness excludes irregularities like sharp turns, providing a foundation for higher-order geometric invariants. A point on the corresponding to t \in I is if the \mathbf{r}'(t) \neq \mathbf{0}, ensuring a well-defined of motion and avoiding self-intersections or stops that could indicate singularities. Conversely, a singular point occurs where \mathbf{r}'(t) = \mathbf{0}, potentially leading to cusps or corners that disrupt the curve's regularity, though such points are excluded in the study of smooth curves. More generally, plane curves can be classified by their differentiability order: a C^k curve is one where the map \mathbf{r} has continuous derivatives up to order k \geq 1, with C^1 requiring only continuous first derivatives for basic tangent analysis. Piecewise smooth curves extend this by allowing a finite number of smooth (C^\infty) segments joined at points where higher derivatives may discontinue, accommodating applications like polygonal approximations while maintaining overall . For a parametrized , the unit tangent vector is given by \mathbf{T}(t) = \frac{\mathbf{r}'(t)}{\|\mathbf{r}'(t)\|}, which provides a normalized direction along the at each point. The term "smooth" in this C^\infty sense was formalized in 20th-century , particularly through Hassler Whitney's work on manifolds, which unified various notions of differentiability into a rigorous framework.

Tangents and Normals

For a plane curve parameterized by a \mathbf{r}(t) = (x(t), y(t)) where t_0 is in the domain and \mathbf{r}'(t_0) \neq \mathbf{0}, the tangent line at the point \mathbf{r}(t_0) is the line passing through that point with direction given by the derivative vector \mathbf{r}'(t_0)./12%3A_Parametric_Equations_and_Polar_Coordinates/12.02%3A_Calculus_of_Parametric_Curves) Its parametric equations are \mathbf{r}(t_0) + s \mathbf{r}'(t_0) for scalar parameter s \in \mathbb{R}./12%3A_Parametric_Equations_and_Polar_Coordinates/12.02%3A_Calculus_of_Parametric_Curves) This line provides the first-order to the curve near t_0, reflecting the instantaneous direction of motion along the curve. For an explicit representation y = f(x) where f is differentiable at x_0, the slope of the tangent line at (x_0, f(x_0)) is f'(x_0), yielding the equation y - f(x_0) = f'(x_0)(x - x_0)./03%3A_Derivatives/3.02%3A_The_Tangent_Line) This form follows directly from the definition of the derivative as the limit of secant slopes./03%3A_Derivatives/3.02%3A_The_Tangent_Line) For an implicit representation defined by f(x, y) = 0 where \nabla f (x_0, y_0) \neq \mathbf{0}, the tangent line at (x_0, y_0) is perpendicular to the gradient vector \nabla f(x_0, y_0) = (f_x(x_0, y_0), f_y(x_0, y_0)). Its equation is \nabla f(x_0, y_0) \cdot (x - x_0, y - y_0) = 0. The gradient points in the direction of steepest ascent on the level set, ensuring the tangent lies in the level set's tangent space. The normal line at a point on the is perpendicular to the line there. For a curve, its direction can be obtained by rotating \mathbf{r}'(t_0) by 90 degrees, such as (-y'(t_0), x'(t_0))./12%3A_Parametric_Equations_and_Polar_Coordinates/12.02%3A_Calculus_of_Parametric_Curves) In the implicit case, the normal line follows the direction, with parametric equations (x_0, y_0) + s \nabla f(x_0, y_0) for s \in \mathbb{R}. The at a point on a smooth plane is the circle that matches both the line and the first-order behavior of the curve's bending at that point, providing a second-order beyond the line alone. At an of a smooth plane , the line intersects the curve with multiplicity at least three, meaning the curve crosses the tangent rather than merely touching it, as the second vanishes while higher derivatives do not. This higher-order contact distinguishes inflections from ordinary points, where the multiplicity is two.

Curvature

The curvature of a smooth plane measures the rate at which the changes along the , quantifying its deviation from a straight line. For a parametrized by s, the \kappa(s) is defined as the magnitude of the of the unit \mathbf{T}(s) with respect to s, i.e., \kappa(s) = \|\mathbf{T}'(s)\|, or equivalently, the rate of change of the \theta(s) that the makes with a fixed , \kappa(s) = \frac{d\theta}{ds}. For a general parametric representation \boldsymbol{\alpha}(t) = (x(t), y(t)) of a plane curve, the curvature is given by \kappa(t) = \frac{|x'(t) y''(t) - y'(t) x''(t)|}{[x'(t)^2 + y'(t)^2]^{3/2}}. This formula arises from the magnitude in the , \kappa(t) = \frac{|\boldsymbol{\alpha}'(t) \times \boldsymbol{\alpha}''(t)|}{\|\boldsymbol{\alpha}'(t)\|^3}, adjusted for the two-dimensional setting. When the curve is expressed explicitly as y = f(x), the curvature simplifies to \kappa(x) = \frac{|f''(x)|}{[1 + f'(x)^2]^{3/2}}. This expression follows directly from substituting the parametric forms x(t) = t and y(t) = f(t) into the general formula. To account for the curve's orientation, the signed curvature is used, defined as \kappa(t) = \frac{x'(t) y''(t) - y'(t) x''(t)}{[x'(t)^2 + y'(t)^2]^{3/2}} for the parametric case, where the sign indicates the direction of turning (positive for counterclockwise rotation relative to the parametrization). The absolute value then yields the unsigned curvature. For a closed smooth plane curve, the total curvature \int \kappa \, ds provides a global measure of bending. Fenchel's theorem states that for any closed curve in the plane (or more generally in space), the integral of the absolute curvature satisfies \int \kappa \, ds \geq 2\pi, with equality the curve is and planar.

Algebraic Plane Curves

Definition and Degree

An algebraic plane curve is defined as the zero set of a single P(x, y) \in k[x, y] of degree d \geq 1 in the affine plane \mathbb{A}^2_k over an k, where irreducibility ensures the curve is not the union of lower-degree curves. The d serves as a primary , classifying curves into lines (d=1), conics (d=2), cubics (d=3), and higher-degree examples, with the polynomial's leading homogeneous part determining asymptotic behavior at infinity. This implicit representation specializes the general framework of plane curves defined by equations, focusing on those arising from irreducible factors. To compactify the affine curve and incorporate points at infinity, the polynomial P(x, y) is homogenized to a homogeneous polynomial P_h(x, y, z) \in k[x, y, z] of the same degree d, yielding the projective closure V(P_h) \subset \mathbb{P}^2_k. Explicitly, if P(x, y) = \sum_{i+j \leq d} a_{i j} x^i y^j, then P_h(x, y, z) = \sum_{i+j \leq d} a_{i j} x^i y^j z^{d - i - j}, which dehomogenizes to P(x, y) upon setting z = 1, ensuring the projective curve contains the original affine curve as an open subset. This construction resolves issues like non-compactness in the affine setting and facilitates the study of global invariants. For smooth projective algebraic plane curves of degree d, the genus g—a topological invariant measuring the number of "holes"—is given by the formula g = \frac{(d-1)(d-2)}{2}. This arises from the adjunction formula relating the canonical divisor to the hyperplane class on \mathbb{P}^2. In particular, for smooth cubics (d=3), g = 1, endowing them with rich arithmetic structure. Curves of genus g=0 are rational, birational to \mathbb{P}^1 and parametrizable by rational functions, while those of genus g=1 are elliptic, admitting a group law via a base point and featuring applications in number theory. Higher genera (g \geq 2) yield more complex curves without rational parametrizations. A fundamental enumerative result is , which states that two curves of degrees d_1 and d_2 with no common irreducible components intersect in exactly d_1 d_2 points, counted with multiplicity. The multiplicity at a point P is the dimension of the quotient of the local ring at P by the ideal generated by the defining s, capturing tangencies and higher-order contacts. This theorem underpins and bounds the number of solutions to systems of equations in the plane.

Singularities and Resolution

In algebraic plane curves, singularities occur at points where the curve fails to be smooth, meaning the partial derivatives of the defining polynomial vanish simultaneously, leading to a breakdown in the usual notions of tangent and regularity. These points are classified based on their local geometry, with nodes representing self-intersections where two branches cross transversally, cusps indicating sharp turns with a single tangent direction but higher-order contact, and ordinary multiple points featuring multiple distinct tangent lines from intersecting branches. The multiplicity m of a singularity at a point P on a curve defined by a polynomial F(x, y) = 0 is the lowest degree of the non-vanishing homogeneous terms in the local Taylor expansion of F around P, equivalently the intersection multiplicity with a general line through P. For instance, ordinary double points have multiplicity 2 and can be nodes or cusps, while higher multiplicities indicate more severe singularities. Resolution of singularities transforms the singular curve into a smooth one via an iterative process of blowing up, where at each step, a singular point is replaced by the of directions through it, pulling back the curve and separating branches until no singularities remain. This birational morphism yields a non-singular model, with the exceptional divisors recording the process; for plane curves, the procedure terminates after finitely many steps due to the dimension. Near a , the local branches can be parameterized using Puiseux series, which are of the form y = a x^{p/q} + higher-order terms in fractional powers, providing a uniformization that resolves the singularity analytically and reveals its topological type. These series arise from the Newton-Puiseux , ensuring every algebraic admits such an in a suitable . A classic example of a node is the curve y^2 = x^2 (x + 1), which has multiplicity 2 at the , where two real branches intersect transversally with distinct tangents along y = \pm x; blowing up once separates them into smooth components. In contrast, the cuspidal cubic y^2 = x^3 exhibits a cusp at the , also of multiplicity 2, with a single tangent y = 0 and intersection multiplicity 3, requiring two blow-ups for resolution, after which the Puiseux expansion y = x^{3/2} parameterizes the branch.

Intersection Properties

The intersection properties of algebraic plane curves are governed by Bézout's theorem, which asserts that two projective plane curves of degrees d_1 and d_2, defined over an algebraically closed field and sharing no common irreducible components, intersect in exactly d_1 d_2 points, counted with multiplicities. The intersection multiplicity I_P(C_1, C_2) at a point P is a local non-negative integer invariant that measures the order of contact between the curves at P; it equals 1 if the curves are smooth at P and their tangent lines differ, and it is greater than 1 otherwise. Specifically, if C_1 and C_2 share the same tangent line at P, then I_P(C_1, C_2) \geq 2, reflecting higher-order tangency. A concrete illustration arises with a line (degree 1) and a conic (degree 2), which intersect at precisely 2 points in the , counting multiplicities; if the line is to the conic at one point, that intersection has multiplicity 2, accounting for the total. In cases where the curves share a common component, the adjusts by considering only proper intersections, leading to fewer distinct points than d_1 d_2. Within linear systems of plane curves—projective spaces parametrizing linear combinations of homogeneous polynomials of fixed —the intersection properties extend to families. Any two distinct members of a complete of dimension at least 1 and d intersect in exactly d^2 points, including fixed base points common to all members and residual intersections that vary with the choice of members. The base locus determines the fixed part of these s, while the residual part captures the moving intersections, with the total multiplicity sum preserved by ; if a common component appears, the residual intersection is obtained by factoring it out from the system. These intersection principles underpin , facilitating counts of curves meeting specified conditions. For instance, exactly one conic passes through five general points in the , since any two distinct conics through those points would intersect at five points, contradicting Bézout's prediction of four intersections unless the conics coincide. Such applications highlight how multiplicity and residual structures resolve apparent overcounts in enumerative problems.

Examples

Conic Sections

Conic sections represent the fundamental examples of algebraic curves of degree two, obtained as intersections of a with a double cone. The general of a conic section in the plane is given by ax^2 + bxy + cy^2 + dx + ey + f = 0, where a, b, c, d, e, f are real coefficients, not all zero. These curves are classified as non-degenerate based on the \delta = b^2 - 4ac: if \delta < 0, the conic is an ellipse (including the circle as a special case when a = c and b = 0); if \delta = 0, it is a parabola; if \delta > 0, it is a . A unifying geometric definition of non-degenerate conic sections is the locus of points P in the such that the of the from P to a fixed point () and the from P to a fixed line (directrix) is a constant e, known as the . For , $0 \leq e < 1; for parabolas, e = 1; for hyperbolas, e > 1. Representative parametric equations include, for an centered at the with semi-major a and semi-minor b (a > b > 0), x = a \cos t, \quad y = b \sin t, \quad t \in [0, 2\pi), and for a parabola such as y^2 = 4ax, x = t^2, \quad y = 2t, \quad t \in \mathbb{R}, which can be scaled appropriately. sections occur when the equation factors into linear terms, resulting in a pair of intersecting straight lines (\delta > 0), parallel lines (\delta = 0), coincident lines, a single point, or the empty set (no real points).

Cubic and Higher Curves

Cubic plane curves are algebraic curves in the defined by a homogeneous polynomial equation of degree three in three variables. Over an , a nonsingular cubic curve is topologically a and has one, making it an in the modern sense. Singular cubics, such as those with a node or cusp, do not possess this structure and have lower genus. Nonsingular cubics can be transformed into Weierstrass form, a standard equation y^2 = x^3 + ax + b, where the \Delta = -16(4a^3 + 27b^2) \neq 0 ensures smoothness; this form facilitates the study of the curve's group law and arithmetic properties. A defining feature of smooth cubics is the presence of exactly nine inflection points, or flexes, which are points where the intersects the curve with multiplicity three; these points play a key role in the curve's embedding and symmetries. For quartics and higher-degree algebraic plane curves, the arithmetic genus of a smooth curve of degree d is given by \frac{(d-1)(d-2)}{2}, yielding genus three for smooth quartics. However, singularities reduce the geometric genus; for instance, a quartic curve with two nodal singularities achieves genus one, resembling an topologically after resolution. Examples include certain spiric sections, intersections of a with a parallel to its , which yield quartic equations. Isomorphism classes of elliptic curves, including those arising from nonsingular cubics, are classified by the , a modular j(E) = 1728 \frac{4a^3}{4a^3 + 27b^2} that remains unchanged under over the base field. This invariant distinguishes distinct elliptic curves up to and connects them to the theory of modular forms.

Transcendental Curves

Transcendental plane curves are those that cannot be defined by algebraic equations involving polynomials and instead require transcendental functions such as exponentials, logarithms, or for their description. Unlike algebraic curves, which possess finite degree and compact projective closures, transcendental curves often exhibit infinite extent and non-compact , extending indefinitely in the . These curves emerged prominently with the of , as their study necessitated and methods inseparable from transcendental expressions. A classic example of a transcendental curve with periodic behavior is the , generated as the roulette trace of a point on the circumference of a rolling along a straight line without slipping. Its parametric equations are given by
x = a(t - \sin t), \quad y = a(1 - \cos t),
where a is the radius of the generating and t is the parameter representing the roll angle. First named by Galileo in after decades of study, the cycloid features repeating arches that extend infinitely along the line of rolling, demonstrating its non-compact nature.
The exemplifies transcendental curves with spiral behavior, defined in polar coordinates by the equation
r = a e^{b \theta},
where r is the radial distance, \theta is the polar angle, a > 0 is a scaling factor, and b determines the growth rate. Invented by Descartes in 1638 and later termed spira mirabilis by Jacob Bernoulli for its self-similar properties, this curve maintains a constant angle between the radius vector and the tangent, known as the equiangular property. Its infinite uncoiling without bound highlights the expansive topology typical of transcendental spirals.
The represents a where an object is dragged along a by a constant-length segment toward a point moving along a straight line, resulting in a with constant length from any point to its . First investigated by Huygens in 1692 and named for its dragging property, it has parametric equations involving , such as
x = a \left( t - \tanh t \right), \quad y = a \sech t,
with x = 0 and finite area between the and , yet infinite length. Its non-compact form tapers asymptotically, illustrating the persistent extension characteristic of many transcendental .
Many transcendental curves, including the examples above, are effectively parameterized using a single variable to capture their evolving shapes over infinite domains.

References

  1. [1]
    [PDF] Plane Algebraic Curves - staff.math.su.se
    Abstract. We go over some of the basics of plane algebraic curves, which are planar curves described as the set of solutions of a polynomial in two ...
  2. [2]
    [PDF] LECTURE NOTES OF MATH 2D 1. Plane curves I – Parametric ...
    What is the meaning of a curve in R2? Intuitionally, this is a subset of R2 with points in it can move in 1 free degree. Examples from pictures: • smooth ...<|control11|><|separator|>
  3. [3]
    Mathematical Treasure: Higher Plane Curves of George Salmon
    In 1852, George Salmon (1819-1904) published his Treatise on Higher Plane Curves. He intended this work as a sequel to his Conic Sections (1848).
  4. [4]
    10.1: Parametrizations of Plane Curves - Mathematics LibreTexts
    Apr 27, 2019 · The graph of parametric equations is called a parametric curve or plane curve, and is denoted by \(C\). Notice in this definition that \(x\) and ...
  5. [5]
    [PDF] Parametric curves in the plane - Penn Math
    This is what we integrate to get arclength of a curve given in parametric form. For example for our circle of radius 2, we have x = 2 cost y = 2 sint and so.
  6. [6]
    [PDF] Curve Representations - Tecgraf/PUC-Rio
    The parametric representation has several advantages. Like the explicit ... equations, parametric equations can also be used to represent closed curves and.
  7. [7]
    Calculus II - Parametric Equations and Curves
    Apr 10, 2025 · In this section we will introduce parametric equations and parametric curves (i.e. graphs of parametric equations). We will graph several ...
  8. [8]
    [PDF] About Implicit Curves in the Plane
    In differential geometry one usually assumes that parametrized curves are without sin- gularities, while in algebraic geometry the singularities of the level ...
  9. [9]
    [PDF] Conversion Methods between Parametric and Implicit Curves and ...
    Oct 17, 1990 · Every plane parametric curve can be expressed as an implicit curve. Some, but not all. implicit curves can be expressed as parametric curves.
  10. [10]
    [PDF] CURVES. 1.1 Parametric and implicit representations.
    A curved C is a plane curve if it is contained in a plane, otherwise it is a space curve. Examples. Determine if the next are plane curves, diferentiable or ...
  11. [11]
    Calculus I - Implicit Differentiation - Pauls Online Math Notes
    Nov 16, 2022 · In implicit differentiation this means that every time we are differentiating a term with y y in it the inside function is the y y and we will ...
  12. [12]
    [PDF] Curves and Surfaces - Carnegie Mellon University
    Feb 19, 2002 · Explicit Representation. • Curve in 2D: y = f(x). • Curve in 3D: y = f(x), z = g(x). • Surface in 3D: z = f(x,y). • Problems: – How about a ...
  13. [13]
    Curves and Surfaces
    An explicit representation of a 2D curve would be represented as y = f(x). An explicit representation of a surface could be represented as z = f(x, y).
  14. [14]
    Tangent and normal lines - Math Insight
    One fundamental interpretation of the derivative of a function is that it is the slope of the tangent line to the graph of the function.
  15. [15]
    The equation of a tangent line
    Use the point-slope formula y−y1=m(x−x1) given m=f′(a) and the point (x1,y1)=(a,f(a)) to get the equation of the line: y−f(a)=f′(a)(x−a).
  16. [16]
    13.3 Arc length and curvature
    The length of the curve y=f(x) between a and b is thus ∫ba√1+(f′(x))2dx. Unfortunately, such integrals are often impossible to do exactly and must be ...
  17. [17]
    [PDF] Curves
    Jan 27, 2014 · definition of orientation of a vector space that do Carmo gives on pages 11 and 12. For a regular curve α the one dimensional vector space ...
  18. [18]
    [PDF] Introduction to Differential Geometry
    Definition. A smooth curve in R3 is a smooth map σ : (a, b) → R3. For each t ∈ (a, b) ...
  19. [19]
    [PDF] DIFFERENTIAL GEOMETRY: A First Course in Curves and Surfaces
    We say f is smooth if f is Ck for every positive integer k. A parametrized curve is a C3 (or smooth) map ˛WI ! R3 for some interval I D .a; b/ or Œa; b ...
  20. [20]
    Hassler Whitney: 1907–1989 - Celebratio Mathematica
    Aug 1, 2013 · By 1930 there were several competing definitions of a differentiable (or smooth) manifold, and in 1932 Veblen and Whitehead had devised an ...
  21. [21]
    Calculus III - Gradient Vector, Tangent Planes and Normal Lines
    Nov 16, 2022 · In this section discuss how the gradient vector can be used to find tangent planes to a much more general function than in the previous ...
  22. [22]
    Osculating Circle -- from Wolfram MathWorld
    The osculating circle of a curve C at a given point P is the circle that has the same tangent as C at point P as well as the same curvature.Missing: source | Show results with:source
  23. [23]
    [PDF] On Inflection Points of Plane Curves
    Since the line is tangent at (0, 1), we know that the intersection multiplicity is strictly larger than one. In this case we get intersection multiplicity 3, ...
  24. [24]
    None
    Below is a merged summary of curvature definitions and Fenchel's Theorem from Do Carmo's *Differential Geometry of Curves and Surfaces*, consolidating all information from the provided segments into a comprehensive and dense format. To maximize detail and clarity, I will use a table in CSV format for the curvature-related information, followed by a detailed narrative for Fenchel's Theorem and additional notes. This approach ensures all unique details are retained while avoiding redundancy.
  25. [25]
    [PDF] DIFFERENTIAL GEOMETRY OF CURVES AND SURFACES 1 ...
    Curvature; the fundamental theorem of plane curves. We would like to measure how “curved” is a curve at a certain point (see Figure 13). In other words, we ...
  26. [26]
    Curvature - Calculus III - Pauls Online Math Notes
    Nov 16, 2022 · There are several formulas for determining the curvature for a curve. The formal definition of curvature is, κ=∥∥∥d→Tds∥∥∥
  27. [27]
    [PDF] curvature.pdf
    To introduce the definition of curvature, we consider a unit-speed curve α(s), where s is the arc length. The tangential angle φ is measured counterclockwise ...
  28. [28]
    2. Curvature and Fenchel's Theorem - Brown Math
    Fenchel also proved the stronger result that the total curvature of a closed curve equals 2 if and only if the curve is a convex plane curve. I. Fáry and J.
  29. [29]
    [PDF] Plane Algebraic Curves - RPTU
    These notes are meant as a gentle introduction to algebraic geometry, a combination of linear algebra and algebra: (a) In linear algebra (as e. g. in the ...
  30. [30]
    [PDF] Lectures on Geometry of Plane Curves An Introduction to Algegraic ...
    Dec 31, 1998 · A plane affine algebraic curve is the locus of a polynomial equation F(X, Y )=0, where F(X, Y ) is a non-constant polynomial.
  31. [31]
    [PDF] Geometry of Algebraic Curves - UChicago Math
    Geometry of Algebraic Curves notes. 3.13 Example. Consider a curve X ⊂ P2, which is a smooth plane curve of degree d. The line bundle O(d) on P2 is also the ...
  32. [32]
    [PDF] Bézout's Theorem for curves - UChicago Math
    Aug 26, 2011 · when we study the points of intersection of two plane curves. ... it counts common points up to multiplicity: the line and the circle intersect.
  33. [33]
    Singularities of plane algebraic curves - ScienceDirect
    Abstract. We give an exposition of some of the basic results on singularities of plane algebraic curves, in terms of polynomials and formal power series.
  34. [34]
    [PDF] ALGEBRAIC CURVES - Mathematics
    Jan 28, 2008 · The intersection number of two plane curves at a point is characterized by its properties, and a definition in terms of a certain residue class ...
  35. [35]
    Plane Algebraic Curves - SpringerLink
    Egbert Brieskorn, Horst Knörrer. Pages 172-202. Definition and elementary properties of plane algebraic curves. Egbert Brieskorn, Horst Knörrer. Pages 202-227 ...
  36. [36]
    [PDF] Bézout's Theorem - IITB Math
    Bézout's theorem relates the number of points at which two algebraic curves intersect to the degrees of the generating polyno- mials. Here we reproduce an ...
  37. [37]
    (PDF) Linear systems of plane curves - ResearchGate
    PDF | On Jan 1, 1999, R. Miranda published Linear systems of plane curves | Find, read and cite all the research you need on ResearchGate.<|control11|><|separator|>
  38. [38]
    [PDF] Enumerative Geometry - RPTU
    Moreover, if we had two distinct smooth conics through the points then these two conics would meet in five points, which is a contradiction to Bézout's theorem.
  39. [39]
    General Form of a Conic | CK-12 Foundation
    The equation B 2 − 4 A C = 0 is used to classify conic sections. If this equation equals zero, the conic section is a parabola.
  40. [40]
    Quadratic Curve -- from Wolfram MathWorld
    The general bivariate quadratic curve can be written ax^2+2bxy+cy^2+2dx+2fy+g=0. Define the following quantities.<|control11|><|separator|>
  41. [41]
    Identifying Conics by the Discriminant (Part 1) - Tree of Math
    Oct 6, 2025 · The expression B2−4AC B 2 − 4 A C is called the discriminant because it helps us discriminate (differentiate) between the different types of ...
  42. [42]
    [PDF] Conic Sections Beyond R2 - Whitman College
    May 14, 2013 · Examples of non-degenerate conics generated by the intersection of a plane ... A conic section can be classified by its discriminant as follows: 1 ...
  43. [43]
    92.18 -- Dissectible cone - UCSB Physics
    One can characterize all conic sections in terms of a focus (fixed point) and a directrix (fixed line), such that the ratio of the distance of a point on the ...
  44. [44]
    Ellipse -- from Wolfram MathWorld
    An ellipse is a curve that is the locus of all points in the plane the sum of whose distances r_1 and r_2 from two fixed points F_1 and F_2 (the foci) ...
  45. [45]
    Parabola Pedal Curve -- from Wolfram MathWorld
    The pedal curve of the parabola with parametric equations x = at^2 (1) y = 2at (2) with pedal point (x_0,y_0) is x_p = ((x_0-a)t^2+y_0t)/(t^2+1) (3) y_p ...
  46. [46]
    [PDF] Elliptic Curves - James Milne
    An elliptic curve is a plane curve defined by a cubic polynomial. Although the problem of finding the rational points on an elliptic curve has fascinated ...<|control11|><|separator|>
  47. [47]
    [PDF] NOTES ON ELLIPTIC CURVES
    We say that a cubic is in (short) Weierstrass form if it is defined by an equation of the form y2 = x3 + ax + b. Proposition 2.4. Let C be a smooth projective ...
  48. [48]
    Obstructions to choosing distinct points on cubic plane curves
    Dec 15, 2018 · Every smooth cubic plane curve has 9 inflection points, 27 sextatic points, and 72 “points of type nine”. Motivated by these classical ...
  49. [49]
    [PDF] Singular quartics of genus one
    g(C) = 3 ∗ 2 2 − 1 − 1=1. Thus a quartic (a curve defined by a degree four polynomial) with two nodal singular- ities is of genus one.
  50. [50]
  51. [51]
    [PDF] The j-invariant of an elliptic curve
    Over an algebraically closed field, two elliptic curves are isomorphic if and only if they have the same j-invariant.
  52. [52]
    Differential Geometry | SpringerLink
    In contrast to algebraic curves, which could be studied in some depth by purely algebraic methods, transcendental curves were inseparable from the methods of ...
  53. [53]
    Differential Geometry | SpringerLink
    Jun 23, 2010 · As mentioned in Chapter 13, calculus made it possible to study nonalgebraic curves: the “mechanical” curves, or transcendental curves as we now ...
  54. [54]
    Differential Geometry
    In contrast to algebraic curves, which could be studied in some depth by purely algebraic methods, transcendental curves were inseparable from the methods of ...
  55. [55]
    Cycloid - MacTutor History of Mathematics
    The cycloid was first studied by Cusa when he was attempting to find the area of a circle by integration. Mersenne gave the first proper definition of the ...
  56. [56]
    Equiangular Spiral - MacTutor History of Mathematics
    The equiangular (or logarithmic) spiral was invented by Descartes in 1638. Torricelli worked on it independently and found the length of the curve.
  57. [57]
    Witch of Agnesi - MacTutor History of Mathematics
    This was studied and named versiera by Maria Agnesi in 1748 in her book Istituzioni Analitiche. It is also known as Cubique d'Agnesi or Agnésienne.
  58. [58]
    Tractrix - MacTutor History of Mathematics - University of St Andrews
    The tractrix is sometimes called a tractory or equitangential curve. It was first studied by Huygens in 1692 who gave it its name.