Plane curve
A plane curve is a curve that lies entirely within a two-dimensional Euclidean plane and can be described either parametrically as the set of points (x(t), y(t)) traced by continuous functions x(t) and y(t) over an interval of the parameter t, or implicitly as the zero locus of a function f(x, y) = 0. Plane curves form a foundational object in both differential and algebraic geometry, enabling the study of geometric properties such as tangents, curvature, and arc length in the parametric case, where the tangent slope at a point is given by \frac{dy/dt}{dx/dt} and the curvature \kappa measures the rate of turning via \kappa = \left| \frac{d\mathbf{T}/dt}{ds/dt} \right| with arc length parameter s.[1] For algebraic plane curves, defined by polynomial equations f(x, y) = 0 of degree d, key invariants include the degree d—with lines at degree 1 and conics at degree 2—and singularities, points where the curve fails to be smooth, affecting properties like the genus.[2] Notable examples of plane curves include straight lines, circles 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 circle, x(t) = a(t - \sin t), y(t) = a(1 - \cos t), which illustrate applications in kinematics. In algebraic geometry, Bézout's theorem states that two plane curves of degrees m and n intersect at exactly mn points in the projective plane, counting multiplicities, underpinning intersection theory.[2] The study of plane curves originated with ancient Greek conic sections and evolved through 19th-century works on higher-degree curves, such as George Salmon's treatise, influencing modern fields like computer graphics and robotics.[3][4]Representations
Parametric Representation
A parametric plane curve is defined as a mapping \mathbf{r}(t) = (x(t), y(t)) from an interval I \subseteq \mathbb{R} to \mathbb{R}^2, where t is the parameter and x(t), y(t) are continuous functions.[5] This representation traces the curve as t varies over I, providing a flexible way to describe paths in the plane.[6] 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 variable.[7] For instance, they symmetrically treat x and y, simplifying computations and extending naturally to complex shapes without restrictions on functionality.[8][6] 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 relation F(x, y) = 0.[5] 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.[8] 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.[6] The orientation of the curve, or the direction of traversal, is determined by the direction in which t increases, such as counterclockwise for the circle example as t rises from 0 to $2\pi.[5]Implicit Representation
An implicit plane curve is defined as the zero set of a function 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 function vanishes, without specifying a parameterization or solving for one variable in terms of the other.[9] Such equations are particularly suited for algebraic manipulations and symmetric descriptions of curves, as they relate x and y directly.[10] The gradient vector \nabla f = \left( \frac{\partial f}{\partial x}, \frac{\partial f}{\partial y} \right) at a point on the curve provides a normal vector to the tangent line, assuming \nabla f \neq (0, 0). This allows implicit determination of the tangent direction as perpendicular to the gradient. Common examples include the unit circle, defined by x^2 + y^2 - 1 = 0, where \nabla f = (2x, 2y), and a straight line, given by ax + by + c = 0, with gradient (a, b) constant along the line.[9][11] To find the slope of the tangent, implicit differentiation 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 function of x along the curve. For the unit circle, this gives \frac{dy}{dx} = -\frac{x}{y}.[12] However, implicit equations can describe curves with multiple connected components, such as disconnected loops, or singular points where the gradient vanishes, complicating analysis. Parameterizing these curves for explicit traversal often requires solving the equation numerically or algebraically.[9][10]Explicit Representation
An explicit plane curve is defined as the graph of a function y = f(x), where f is a real-valued function defined on some interval or domain in the real numbers, representing the set of points (x, f(x)) in the Cartesian plane.[13] 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.[14] However, this representation has notable limitations. It cannot capture curves with vertical tangents, such as a vertical line x = c, because no function y = f(x) can assign multiple y-values to a single x or handle undefined slopes at vertical points.[13] Additionally, it fails to represent closed loops, like a circle, or multi-branched curves, where a single x might correspond to multiple y-values, requiring piecewise definitions or alternative approaches.[13] 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.[15] For instance, the equation of the tangent line at x = a is y - f(a) = f'(a)(x - a).[16] 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.[17] 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.[13]Smooth Plane Curves
Definition and Differentiability
In differential geometry, 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.[18] 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.[19] This C^\infty smoothness excludes irregularities like sharp turns, providing a foundation for higher-order geometric invariants.[20] A point on the curve corresponding to parameter t \in I is regular if the derivative \mathbf{r}'(t) \neq \mathbf{0}, ensuring a well-defined direction of motion and avoiding self-intersections or stops that could indicate singularities.[18] 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.[19] 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.[20] 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 continuity.[20] For a regular parametrized curve, the unit tangent vector is given by \mathbf{T}(t) = \frac{\mathbf{r}'(t)}{\|\mathbf{r}'(t)\|}, which provides a normalized direction along the curve at each regular point.[19] The term "smooth" in this C^\infty sense was formalized in 20th-century differential geometry, particularly through Hassler Whitney's 1936 work on manifolds, which unified various notions of differentiability into a rigorous framework.[21]Tangents and Normals
For a smooth plane curve parameterized by a vector-valued function \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 linear approximation 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)).[22] Its equation is \nabla f(x_0, y_0) \cdot (x - x_0, y - y_0) = 0.[22] The gradient points in the direction of steepest ascent on the level set, ensuring the tangent lies in the level set's tangent space.[22] The normal line at a point on the curve is perpendicular to the tangent line there. For a parametric 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 gradient direction, with parametric equations (x_0, y_0) + s \nabla f(x_0, y_0) for s \in \mathbb{R}.[22] The osculating circle at a point on a smooth plane curve is the circle that matches both the tangent line and the first-order behavior of the curve's bending at that point, providing a second-order approximation beyond the tangent line alone.[23] At an inflection point of a smooth plane curve, the tangent line intersects the curve with multiplicity at least three, meaning the curve crosses the tangent rather than merely touching it, as the second derivative vanishes while higher derivatives do not.[24] This higher-order contact distinguishes inflections from ordinary points, where the multiplicity is two.[24]Curvature
The curvature of a smooth plane curve measures the rate at which the tangent direction changes along the curve, quantifying its deviation from a straight line. For a curve parametrized by arc length s, the curvature \kappa(s) is defined as the magnitude of the derivative of the unit tangent vector \mathbf{T}(s) with respect to s, i.e., \kappa(s) = \|\mathbf{T}'(s)\|, or equivalently, the rate of change of the angle \theta(s) that the tangent makes with a fixed direction, \kappa(s) = \frac{d\theta}{ds}.[25] 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 cross product magnitude in the plane, \kappa(t) = \frac{|\boldsymbol{\alpha}'(t) \times \boldsymbol{\alpha}''(t)|}{\|\boldsymbol{\alpha}'(t)\|^3}, adjusted for the two-dimensional setting.[25][26] 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.[25][27] 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.[25][28] 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 if and only if the curve is convex and planar.[25][29]Algebraic Plane Curves
Definition and Degree
An algebraic plane curve is defined as the zero set of a single irreducible polynomial P(x, y) \in k[x, y] of degree d \geq 1 in the affine plane \mathbb{A}^2_k over an algebraically closed field k, where irreducibility ensures the curve is not the union of lower-degree curves.[30] The degree d serves as a primary invariant, 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.[30] This implicit representation specializes the general framework of plane curves defined by polynomial equations, focusing on those arising from irreducible factors.[31] 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.[30] 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.[31] This construction resolves issues like non-compactness in the affine setting and facilitates the study of global invariants.[30] 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.[32] In particular, for smooth cubics (d=3), g = 1, endowing them with rich arithmetic structure.[32] 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.[32] Higher genera (g \geq 2) yield more complex curves without rational parametrizations.[32] A fundamental enumerative result is Bézout's theorem, which states that two projective plane curves of degrees d_1 and d_2 with no common irreducible components intersect in exactly d_1 d_2 points, counted with multiplicity.[33] The intersection multiplicity at a point P is the dimension of the quotient of the local ring at P by the ideal generated by the defining polynomials, capturing tangencies and higher-order contacts.[33] This theorem underpins intersection theory and bounds the number of solutions to systems of polynomial equations in the plane.[33]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.[34][35] 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.[34][35][36] 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 projective line 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 resolution process; for plane curves, the procedure terminates after finitely many steps due to the dimension.[34][35][36] Near a singularity, the local branches can be parameterized using Puiseux series, which are expansions 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 theorem, ensuring every algebraic branch admits such an expansion in a suitable coordinate system.[34][36][35] A classic example of a node is the curve y^2 = x^2 (x + 1), which has multiplicity 2 at the origin, 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 origin, 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.[34][35]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.[30] 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.[30] 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.[33] A concrete illustration arises with a line (degree 1) and a conic (degree 2), which intersect at precisely 2 points in the projective plane, counting multiplicities; if the line is tangent to the conic at one point, that intersection has multiplicity 2, accounting for the total.[37] In cases where the curves share a common component, the theorem adjusts by considering only proper intersections, leading to fewer distinct points than d_1 d_2.[30] Within linear systems of plane curves—projective spaces parametrizing linear combinations of homogeneous polynomials of fixed degree—the intersection properties extend to families. Any two distinct members of a complete linear system of dimension at least 1 and degree 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.[30] The base locus determines the fixed part of these intersections, while the residual part captures the moving intersections, with the total multiplicity sum preserved by Bézout's theorem; if a common component appears, the residual intersection is obtained by factoring it out from the system.[38] These intersection principles underpin enumerative geometry, facilitating counts of curves meeting specified conditions. For instance, exactly one conic passes through five general points in the projective plane, since any two distinct conics through those points would intersect at five points, contradicting Bézout's prediction of four intersections unless the conics coincide.[39] Such applications highlight how multiplicity and residual structures resolve apparent overcounts in enumerative problems.[39]Examples
Conic Sections
Conic sections represent the fundamental examples of algebraic plane curves of degree two, obtained as intersections of a plane with a double cone. The general equation 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.[40] These curves are classified as non-degenerate based on the discriminant \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 hyperbola.[41][42] A unifying geometric definition of non-degenerate conic sections is the locus of points P in the plane such that the ratio of the distance from P to a fixed point (focus) and the distance from P to a fixed line (directrix) is a constant e, known as the eccentricity. For ellipses, $0 \leq e < 1; for parabolas, e = 1; for hyperbolas, e > 1.[43][44][45] Representative parametric equations include, for an ellipse centered at the origin with semi-major axis a and semi-minor axis 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.[45][46] Degenerate conic 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).[41]Cubic and Higher Curves
Cubic plane curves are algebraic curves in the projective plane defined by a homogeneous polynomial equation of degree three in three variables. Over an algebraically closed field, a nonsingular cubic curve is topologically a torus and has genus one, making it an elliptic curve in the modern sense.[47] 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 discriminant \Delta = -16(4a^3 + 27b^2) \neq 0 ensures smoothness; this form facilitates the study of the curve's group law and arithmetic properties.[48] A defining feature of smooth cubics is the presence of exactly nine inflection points, or flexes, which are points where the tangent intersects the curve with multiplicity three; these points play a key role in the curve's embedding and symmetries.[49] 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 elliptic curve topologically after resolution.[50] Examples include certain spiric sections, intersections of a torus with a plane parallel to its axis, which yield quartic equations.[51] Isomorphism classes of elliptic curves, including those arising from nonsingular cubics, are classified by the j-invariant, a modular function j(E) = 1728 \frac{4a^3}{4a^3 + 27b^2} that remains unchanged under isomorphism over the base field.[52] This invariant distinguishes distinct elliptic curves up to isomorphism and connects them to the theory of modular forms.[47]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 trigonometric functions for their description.[53] Unlike algebraic curves, which possess finite degree and compact projective closures, transcendental curves often exhibit infinite extent and non-compact topology, extending indefinitely in the plane.[54] These curves emerged prominently with the development of calculus, as their study necessitated differential and integral methods inseparable from transcendental expressions.[55] A classic example of a transcendental curve with periodic behavior is the cycloid, generated as the roulette trace of a point on the circumference of a circle rolling along a straight line without slipping.[56] Its parametric equations are given byx = a(t - \sin t), \quad y = a(1 - \cos t),
where a is the radius of the generating circle and t is the parameter representing the roll angle.[56] First named by Galileo in 1599 after decades of study, the cycloid features repeating arches that extend infinitely along the line of rolling, demonstrating its non-compact nature.[56] The logarithmic spiral 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.[57] 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.[57] Its infinite uncoiling without bound highlights the expansive topology typical of transcendental spirals.[57] The tractrix represents a pursuit curve where an object is dragged along a plane by a constant-length tangent segment toward a point moving along a straight line, resulting in a curve with constant tangent length from any point to its asymptote.[58] First investigated by Huygens in 1692 and named tractrix for its dragging property, it has parametric equations involving hyperbolic functions, such as
x = a \left( t - \tanh t \right), \quad y = a \sech t,
with asymptote x = 0 and finite area between the curve and asymptote, yet infinite length.[58] Its non-compact form tapers asymptotically, illustrating the persistent extension characteristic of many transcendental curves.[58] Many transcendental curves, including the examples above, are effectively parameterized using a single variable to capture their evolving shapes over infinite domains.[53]