Arc length

Last updated
When rectified, the curve gives a straight line segment with the same length as the curve's arc length. Arc length.gif
When rectified, the curve gives a straight line segment with the same length as the curve's arc length.
Arc length s of a logarithmic spiral as a function of its parameter th. Logarithmic spiral arc length.gif
Arc length s of a logarithmic spiral as a function of its parameter θ.

Arc length is the distance between two points along a section of a curve.

Contents

Determining the length of an irregular arc segment by approximating the arc segment as connected (straight) line segments is also called curve rectification. A rectifiable curve has a finite number of segments in its rectification (so the curve has a finite length).

If a curve can be parameterized as an injective and continuously differentiable function (i.e., the derivative is a continuous function) , then the curve is rectifiable (i.e., it has a finite length).

The advent of infinitesimal calculus led to a general formula that provides closed-form solutions in some cases.

General approach

Approximation to a curve by multiple linear segments, called rectification of a curve. Arclength.svg
Approximation to a curve by multiple linear segments, called rectification of a curve.

A curve in the plane can be approximated by connecting a finite number of points on the curve using (straight) line segments to create a polygonal path. Since it is straightforward to calculate the length of each linear segment (using the Pythagorean theorem in Euclidean space, for example), the total length of the approximation can be found by summation of the lengths of each linear segment; that approximation is known as the (cumulative) chordal distance. [1]

If the curve is not already a polygonal path, then using a progressively larger number of line segments of smaller lengths will result in better curve length approximations. Such a curve length determination by approximating the curve as connected (straight) line segments is called rectification of a curve. The lengths of the successive approximations will not decrease and may keep increasing indefinitely, but for smooth curves they will tend to a finite limit as the lengths of the segments get arbitrarily small.

For some curves, there is a smallest number that is an upper bound on the length of all polygonal approximations (rectification). These curves are called rectifiable and the arc length is defined as the number .

A signed arc length can be defined to convey a sense of orientation or "direction" with respect to a reference point taken as origin in the curve (see also: curve orientation and signed distance). [2]

Formula for a smooth curve

Let be an injective and continuously differentiable (i.e., the derivative is a continuous function) function. The length of the curve defined by can be defined as the limit of the sum of linear segment lengths for a regular partition of as the number of segments approaches infinity. This means

where with for This definition is equivalent to the standard definition of arc length as an integral:

The last equality is proved by the following steps:

  1. The second fundamental theorem of calculus shows
    where over maps to and . In the below step, the following equivalent expression is used.
  2. The function is a continuous function from a closed interval to the set of real numbers, thus it is uniformly continuous according to the Heine–Cantor theorem, so there is a positive real and monotonically non-decreasing function of positive real numbers such that implies where and . Let's consider the limit of the following formula,

With the above step result, it becomes

Terms are rearranged so that it becomes

where in the leftmost side is used. By for so that , it becomes

with , , and . In the limit so thus the left side of approaches . In other words, in this limit, and the right side of this equality is just the Riemann integral of on This definition of arc length shows that the length of a curve represented by a continuously differentiable function on is always finite, i.e., rectifiable.

The definition of arc length of a smooth curve as the integral of the norm of the derivative is equivalent to the definition

where the supremum is taken over all possible partitions of [3] This definition as the supremum of the all possible partition sums is also valid if is merely continuous, not differentiable.

A curve can be parameterized in infinitely many ways. Let be any continuously differentiable bijection. Then is another continuously differentiable parameterization of the curve originally defined by The arc length of the curve is the same regardless of the parameterization used to define the curve:

Finding arc lengths by integration

Quarter circle Quarter circle.png
Quarter circle

If a planar curve in is defined by the equation where is continuously differentiable, then it is simply a special case of a parametric equation where and The Euclidean distance of each infinitesimal segment of the arc can be given by:

The arc length is then given by:

Curves with closed-form solutions for arc length include the catenary, circle, cycloid, logarithmic spiral, parabola, semicubical parabola and straight line. The lack of a closed form solution for the arc length of an elliptic and hyperbolic arc led to the development of the elliptic integrals.

Numerical integration

In most cases, including even simple curves, there are no closed-form solutions for arc length and numerical integration is necessary. Numerical integration of the arc length integral is usually very efficient. For example, consider the problem of finding the length of a quarter of the unit circle by numerically integrating the arc length integral. The upper half of the unit circle can be parameterized as The interval corresponds to a quarter of the circle. Since and the length of a quarter of the unit circle is

The 15-point Gauss–Kronrod rule estimate for this integral of 1.570796326808177 differs from the true length of

by 1.3×10−11 and the 16-point Gaussian quadrature rule estimate of 1.570796326794727 differs from the true length by only 1.7×10−13. This means it is possible to evaluate this integral to almost machine precision with only 16 integrand evaluations.

Curve on a surface

Let be a surface mapping and let be a curve on this surface. The integrand of the arc length integral is Evaluating the derivative requires the chain rule for vector fields:

The squared norm of this vector is

(where is the first fundamental form coefficient), so the integrand of the arc length integral can be written as (where and ).

Other coordinate systems

Let be a curve expressed in polar coordinates. The mapping that transforms from polar coordinates to rectangular coordinates is

The integrand of the arc length integral is The chain rule for vector fields shows that So the squared integrand of the arc length integral is

So for a curve expressed in polar coordinates, the arc length is:

The second expression is for a polar graph parameterized by .

Now let be a curve expressed in spherical coordinates where is the polar angle measured from the positive -axis and is the azimuthal angle. The mapping that transforms from spherical coordinates to rectangular coordinates is

Using the chain rule again shows that All dot products where and differ are zero, so the squared norm of this vector is

So for a curve expressed in spherical coordinates, the arc length is

A very similar calculation shows that the arc length of a curve expressed in cylindrical coordinates is

Simple cases

Arcs of circles

Arc lengths are denoted by s, since the Latin word for length (or size) is spatium.

In the following lines, represents the radius of a circle, is its diameter, is its circumference, is the length of an arc of the circle, and is the angle which the arc subtends at the centre of the circle. The distances and are expressed in the same units.

Great circles on Earth

Two units of length, the nautical mile and the metre (or kilometre), were originally defined so the lengths of arcs of great circles on the Earth's surface would be simply numerically related to the angles they subtend at its centre. The simple equation applies in the following circumstances:

  • if is in nautical miles, and is in arcminutes (160 degree), or
  • if is in kilometres, and is in gradians.

The lengths of the distance units were chosen to make the circumference of the Earth equal 40000 kilometres, or 21600 nautical miles. Those are the numbers of the corresponding angle units in one complete turn.

Those definitions of the metre and the nautical mile have been superseded by more precise ones, but the original definitions are still accurate enough for conceptual purposes and some calculations. For example, they imply that one kilometre is exactly 0.54 nautical miles. Using official modern definitions, one nautical mile is exactly 1.852 kilometres, [4] which implies that 1 kilometre is about 0.53995680 nautical miles. [5] This modern ratio differs from the one calculated from the original definitions by less than one part in 10,000.

Other simple cases

Historical methods

Antiquity

For much of the history of mathematics, even the greatest thinkers considered it impossible to compute the length of an irregular arc. Although Archimedes had pioneered a way of finding the area beneath a curve with his "method of exhaustion", few believed it was even possible for curves to have definite lengths, as do straight lines. The first ground was broken in this field, as it often has been in calculus, by approximation. People began to inscribe polygons within the curves and compute the length of the sides for a somewhat accurate measurement of the length. By using more segments, and by decreasing the length of each segment, they were able to obtain a more and more accurate approximation. In particular, by inscribing a polygon of many sides in a circle, they were able to find approximate values of π. [6] [7]

17th century

In the 17th century, the method of exhaustion led to the rectification by geometrical methods of several transcendental curves: the logarithmic spiral by Evangelista Torricelli in 1645 (some sources say John Wallis in the 1650s), the cycloid by Christopher Wren in 1658, and the catenary by Gottfried Leibniz in 1691.

In 1659, Wallis credited William Neile's discovery of the first rectification of a nontrivial algebraic curve, the semicubical parabola. [8] The accompanying figures appear on page 145. On page 91, William Neile is mentioned as Gulielmus Nelius.

Integral form

Before the full formal development of calculus, the basis for the modern integral form for arc length was independently discovered by Hendrik van Heuraet and Pierre de Fermat.

In 1659 van Heuraet published a construction showing that the problem of determining arc length could be transformed into the problem of determining the area under a curve (i.e., an integral). As an example of his method, he determined the arc length of a semicubical parabola, which required finding the area under a parabola. [9] In 1660, Fermat published a more general theory containing the same result in his De linearum curvarum cum lineis rectis comparatione dissertatio geometrica (Geometric dissertation on curved lines in comparison with straight lines). [10]

Fermat's method of determining arc length Arc length, Fermat.svg
Fermat's method of determining arc length

Building on his previous work with tangents, Fermat used the curve

whose tangent at x = a had a slope of

so the tangent line would have the equation

Next, he increased a by a small amount to a + ε, making segment AC a relatively good approximation for the length of the curve from A to D. To find the length of the segment AC, he used the Pythagorean theorem:

which, when solved, yields

In order to approximate the length, Fermat would sum up a sequence of short segments.

Curves with infinite length

The Koch curve. Koch curve.svg
The Koch curve.
The graph of
x
[?]
sin
[?]
(
1
/
x
)
{\displaystyle x\cdot \sin(1/x)} Xsinoneoverx.svg
The graph of

As mentioned above, some curves are non-rectifiable. That is, there is no upper bound on the lengths of polygonal approximations; the length can be made arbitrarily large. Informally, such curves are said to have infinite length. There are continuous curves on which every arc (other than a single-point arc) has infinite length. An example of such a curve is the Koch curve. Another example of a curve with infinite length is the graph of the function defined by f(x) = x sin(1/x) for any open set with 0 as one of its delimiters and f(0) = 0. Sometimes the Hausdorff dimension and Hausdorff measure are used to quantify the size of such curves.

Generalization to (pseudo-)Riemannian manifolds

Let be a (pseudo-)Riemannian manifold, the (pseudo-) metric tensor, a curve in defined by parametric equations

and

The length of , is defined to be

,

or, choosing local coordinates ,

,

where

is the tangent vector of at The sign in the square root is chosen once for a given curve, to ensure that the square root is a real number. The positive sign is chosen for spacelike curves; in a pseudo-Riemannian manifold, the negative sign may be chosen for timelike curves. Thus the length of a curve is a non-negative real number. Usually no curves are considered which are partly spacelike and partly timelike.

In theory of relativity, arc length of timelike curves (world lines) is the proper time elapsed along the world line, and arc length of a spacelike curve the proper distance along the curve.

See also

Related Research Articles

<span class="mw-page-title-main">Centripetal force</span> Force directed to the center of rotation

A centripetal force is a force that makes a body follow a curved path. The direction of the centripetal force is always orthogonal to the motion of the body and towards the fixed point of the instantaneous center of curvature of the path. Isaac Newton described it as "a force by which bodies are drawn or impelled, or in any way tend, towards a point as to a centre". In the theory of Newtonian mechanics, gravity provides the centripetal force causing astronomical orbits.

<span class="mw-page-title-main">Dirac delta function</span> Generalized function whose value is zero everywhere except at zero

In mathematical physics, the Dirac delta distribution, also known as the unit impulse, is a generalized function or distribution over the real numbers, whose value is zero everywhere except at zero, and whose integral over the entire real line is equal to one.

<span class="mw-page-title-main">Equations of motion</span> Equations that describe the behavior of a physical system

In physics, equations of motion are equations that describe the behavior of a physical system in terms of its motion as a function of time. More specifically, the equations of motion describe the behavior of a physical system as a set of mathematical functions in terms of dynamic variables. These variables are usually spatial coordinates and time, but may include momentum components. The most general choice are generalized coordinates which can be any convenient variables characteristic of the physical system. The functions are defined in a Euclidean space in classical mechanics, but are replaced by curved spaces in relativity. If the dynamics of a system is known, the equations are the solutions for the differential equations describing the motion of the dynamics.

<span class="mw-page-title-main">Work (physics)</span> Process of energy transfer to an object via force application through displacement

In physics, work is the energy transferred to or from an object via the application of force along a displacement. In its simplest form, for a constant force aligned with the direction of motion, the work equals the product of the force strength and the distance traveled. A force is said to do positive work if when applied it has a component in the direction of the displacement of the point of application. A force does negative work if it has a component opposite to the direction of the displacement at the point of application of the force.

<span class="mw-page-title-main">Noether's theorem</span> Statement relating differentiable symmetries to conserved quantities

Noether's theorem or Noether's first theorem states that every differentiable symmetry of the action of a physical system with conservative forces has a corresponding conservation law. The theorem was proven by mathematician Emmy Noether in 1915 and published in 1918. The action of a physical system is the integral over time of a Lagrangian function, from which the system's behavior can be determined by the principle of least action. This theorem only applies to continuous and smooth symmetries over physical space.

<span class="mw-page-title-main">Laplace operator</span> Differential operator

In mathematics, the Laplace operator or Laplacian is a differential operator given by the divergence of the gradient of a scalar function on Euclidean space. It is usually denoted by the symbols , (where is the nabla operator), or . In a Cartesian coordinate system, the Laplacian is given by the sum of second partial derivatives of the function with respect to each independent variable. In other coordinate systems, such as cylindrical and spherical coordinates, the Laplacian also has a useful form. Informally, the Laplacian Δf (p) of a function f at a point p measures by how much the average value of f over small spheres or balls centered at p deviates from f (p).

In continuum mechanics, the infinitesimal strain theory is a mathematical approach to the description of the deformation of a solid body in which the displacements of the material particles are assumed to be much smaller than any relevant dimension of the body; so that its geometry and the constitutive properties of the material at each point of space can be assumed to be unchanged by the deformation.

<span class="mw-page-title-main">Green's theorem</span> Theorem in calculus relating line and double integrals

In vector calculus, Green's theorem relates a line integral around a simple closed curve C to a double integral over the plane region D bounded by C. It is the two-dimensional special case of Stokes' theorem.

Linear elasticity is a mathematical model of how solid objects deform and become internally stressed due to prescribed loading conditions. It is a simplification of the more general nonlinear theory of elasticity and a branch of continuum mechanics.

In analytical mechanics, generalized coordinates are a set of parameters used to represent the state of a system in a configuration space. These parameters must uniquely define the configuration of the system relative to a reference state. The generalized velocities are the time derivatives of the generalized coordinates of the system. The adjective "generalized" distinguishes these parameters from the traditional use of the term "coordinate" to refer to Cartesian coordinates.

<span class="mw-page-title-main">Directional derivative</span> Instantaneous rate of change of the function

A directional derivative is a concept in multivariable calculus that measures the rate at which a function changes in a particular direction at a given point.

<span class="mw-page-title-main">Envelope (mathematics)</span> Family of curves in geometry

In geometry, an envelope of a planar family of curves is a curve that is tangent to each member of the family at some point, and these points of tangency together form the whole envelope. Classically, a point on the envelope can be thought of as the intersection of two "infinitesimally adjacent" curves, meaning the limit of intersections of nearby curves. This idea can be generalized to an envelope of surfaces in space, and so on to higher dimensions.

<span class="mw-page-title-main">Hamilton–Jacobi equation</span> A reformulation of Newtons laws of motion using the calculus of variations

In physics, the Hamilton–Jacobi equation, named after William Rowan Hamilton and Carl Gustav Jacob Jacobi, is an alternative formulation of classical mechanics, equivalent to other formulations such as Newton's laws of motion, Lagrangian mechanics and Hamiltonian mechanics.

In mechanics, virtual work arises in the application of the principle of least action to the study of forces and movement of a mechanical system. The work of a force acting on a particle as it moves along a displacement is different for different displacements. Among all the possible displacements that a particle may follow, called virtual displacements, one will minimize the action. This displacement is therefore the displacement followed by the particle according to the principle of least action.

The work of a force on a particle along a virtual displacement is known as the virtual work.

<span class="mw-page-title-main">Osculating circle</span> Circle of immediate corresponding curvature of a curve at a point

In differential geometry of curves, the osculating circle of a sufficiently smooth plane curve at a given point p on the curve has been traditionally defined as the circle passing through p and a pair of additional points on the curve infinitesimally close to p. Its center lies on the inner normal line, and its curvature defines the curvature of the given curve at that point. This circle, which is the one among all tangent circles at the given point that approaches the curve most tightly, was named circulus osculans by Leibniz.

<span class="mw-page-title-main">Leibniz integral rule</span> Differentiation under the integral sign formula

In calculus, the Leibniz integral rule for differentiation under the integral sign states that for an integral of the form

<span class="mw-page-title-main">Hamilton's principle</span> Formulation of the principle of stationary action

In physics, Hamilton's principle is William Rowan Hamilton's formulation of the principle of stationary action. It states that the dynamics of a physical system are determined by a variational problem for a functional based on a single function, the Lagrangian, which may contain all physical information concerning the system and the forces acting on it. The variational problem is equivalent to and allows for the derivation of the differential equations of motion of the physical system. Although formulated originally for classical mechanics, Hamilton's principle also applies to classical fields such as the electromagnetic and gravitational fields, and plays an important role in quantum mechanics, quantum field theory and criticality theories.

<span class="mw-page-title-main">Pendulum (mechanics)</span> Free swinging suspended body

A pendulum is a body suspended from a fixed support so that it swings freely back and forth under the influence of gravity. When a pendulum is displaced sideways from its resting, equilibrium position, it is subject to a restoring force due to gravity that will accelerate it back toward the equilibrium position. When released, the restoring force acting on the pendulum's mass causes it to oscillate about the equilibrium position, swinging it back and forth. The mathematics of pendulums are in general quite complicated. Simplifying assumptions can be made, which in the case of a simple pendulum allow the equations of motion to be solved analytically for small-angle oscillations.

<span class="mw-page-title-main">Line integral</span> Definite integral of a scalar or vector field along a path

In mathematics, a line integral is an integral where the function to be integrated is evaluated along a curve. The terms path integral, curve integral, and curvilinear integral are also used; contour integral is used as well, although that is typically reserved for line integrals in the complex plane.

<span class="mw-page-title-main">Lagrangian mechanics</span> Formulation of classical mechanics

In physics, Lagrangian mechanics is a formulation of classical mechanics founded on the stationary-action principle. It was introduced by the Italian-French mathematician and astronomer Joseph-Louis Lagrange in his 1788 work, Mécanique analytique.

References

  1. Ahlberg; Nilson (1967). The Theory of Splines and Their Applications . Academic Press. p.  51. ISBN   9780080955452.
  2. Nestoridis, Vassili; Papadopoulos, Athanase (2017). "Arc length as a global conformal parameter for analytic curves". Journal of Mathematical Analysis and Applications. Elsevier BV. 445 (2): 1505–1515. doi: 10.1016/j.jmaa.2016.02.031 . ISSN   0022-247X.
  3. Rudin, Walter (1976). Principles of Mathematical Analysis . McGraw-Hill, Inc. pp.  137. ISBN   978-0-07-054235-8.
  4. Suplee, Curt (2 July 2009). "Special Publication 811". nist.gov.
  5. CRC Handbook of Chemistry and Physics, p. F-254
  6. Richeson, David (May 2015). "Circular Reasoning: Who First Proved That C Divided by d Is a Constant?". The College Mathematics Journal. 46 (3): 162–171. doi:10.4169/college.math.j.46.3.162. ISSN   0746-8342. S2CID   123757069.
  7. Coolidge, J. L. (February 1953). "The Lengths of Curves". The American Mathematical Monthly. 60 (2): 89–93. doi:10.2307/2308256. JSTOR   2308256.
  8. Wallis, John (1659). Tractatus Duo. Prior, De Cycloide et de Corporibus inde Genitis…. Oxford: University Press. pp. 91–96.
  9. van Heuraet, Hendrik (1659). "Epistola de transmutatione curvarum linearum in rectas [Letter on the transformation of curved lines into right ones]". Renati Des-Cartes Geometria (2nd ed.). Amsterdam: Louis & Daniel Elzevir. pp. 517–520.
  10. M.P.E.A.S. (pseudonym of Fermat) (1660). De Linearum Curvarum cum Lineis Rectis Comparatione Dissertatio Geometrica. Toulouse: Arnaud Colomer.

Sources