Spherical trigonometry

Last updated

The octant of a sphere is a spherical triangle with three right angles. Triangle trirectangle.png
The octant of a sphere is a spherical triangle with three right angles.

Spherical trigonometry is the branch of spherical geometry that deals with the metrical relationships between the sides and angles of spherical triangles, traditionally expressed using trigonometric functions. On the sphere, geodesics are great circles. Spherical trigonometry is of great importance for calculations in astronomy, geodesy, and navigation.

Contents

The origins of spherical trigonometry in Greek mathematics and the major developments in Islamic mathematics are discussed fully in History of trigonometry and Mathematics in medieval Islam. The subject came to fruition in Early Modern times with important developments by John Napier, Delambre and others, and attained an essentially complete form by the end of the nineteenth century with the publication of Todhunter's textbook Spherical trigonometry for the use of colleges and Schools. [1] Since then, significant developments have been the application of vector methods, quaternion methods, and the use of numerical methods.

Preliminaries

Eight spherical triangles defined by the intersection of three great circles. Spherical trigonometry Intersecting circles.svg
Eight spherical triangles defined by the intersection of three great circles.

Spherical polygons

A spherical polygon is a polygon on the surface of the sphere. Its sides are arcs of great circles—the spherical geometry equivalent of line segments in plane geometry.

Such polygons may have any number of sides greater than 1. Two-sided spherical polygons— lunes , also called digons or bi-angles—are bounded by two great-circle arcs: a familiar example is the curved outward-facing surface of a segment of an orange. Three arcs serve to define a spherical triangle, the principal subject of this article. Polygons with higher numbers of sides (4-sided spherical quadrilaterals, 5-sided spherical pentagons, etc.) are defined in similar manner. Analogously to their plane counterparts, spherical polygons with more than 3 sides can always be treated as the composition of spherical triangles.

One spherical polygon with interesting properties is the pentagramma mirificum, a 5-sided spherical star polygon with a right angle at every vertex.

From this point in the article, discussion will be restricted to spherical triangles, referred to simply as triangles.

Notation

The basic triangle on a unit sphere. Spherical trigonometry basic triangle.svg
The basic triangle on a unit sphere.

In particular, the sum of the angles of a spherical triangle is strictly greater than the sum of the angles of a triangle defined on the Euclidean plane, which is always exactly π radians.

Polar triangles

The polar triangle ^A'B'C' Spherical trigonometry polar triangle.svg
The polar triangle A'B'C'

The polar triangle associated with a triangle ABC is defined as follows. Consider the great circle that contains the side BC. This great circle is defined by the intersection of a diametral plane with the surface. Draw the normal to that plane at the centre: it intersects the surface at two points and the point that is on the same side of the plane as A is (conventionally) termed the pole of A and it is denoted by A'. The points B' and C' are defined similarly.

The triangle A'B'C' is the polar triangle corresponding to triangle ABC. A very important theorem (Todhunter, [1] Art.27) proves that the angles and sides of the polar triangle are given by

Therefore, if any identity is proved for ABC then we can immediately derive a second identity by applying the first identity to the polar triangle by making the above substitutions. This is how the supplemental cosine equations are derived from the cosine equations. Similarly, the identities for a quadrantal triangle can be derived from those for a right-angled triangle. The polar triangle of a polar triangle is the original triangle.

Cosine rules and sine rules

Cosine rules

The cosine rule is the fundamental identity of spherical trigonometry: all other identities, including the sine rule, may be derived from the cosine rule:

These identities generalize the cosine rule of plane trigonometry, to which they are asymptotically equivalent in the limit of small interior angles. (On the unit sphere, if set and etc.; see Spherical law of cosines.)

Sine rules

The spherical law of sines is given by the formula

These identities approximate the sine rule of plane trigonometry when the sides are much smaller than the radius of the sphere.

Derivation of the cosine rule

Spherical trigonometry vectors.svg

The spherical cosine formulae were originally proved by elementary geometry and the planar cosine rule (Todhunter, [1] Art.37). He also gives a derivation using simple coordinate geometry and the planar cosine rule (Art.60). The approach outlined here uses simpler vector methods. (These methods are also discussed at Spherical law of cosines.)

Consider three unit vectors OA, OB, OC drawn from the origin to the vertices of the triangle (on the unit sphere). The arc BC subtends an angle of magnitude a at the centre and therefore OB · OC = cos a. Introduce a Cartesian basis with OA along the z-axis and OB in the xz-plane making an angle c with the z-axis. The vector OC projects to ON in the xy-plane and the angle between ON and the x-axis is A. Therefore, the three vectors have components:

The scalar product OB · OC in terms of the components is

Equating the two expressions for the scalar product gives

This equation can be re-arranged to give explicit expressions for the angle in terms of the sides:

The other cosine rules are obtained by cyclic permutations.

Derivation of the sine rule

This derivation is given in Todhunter, [1] (Art.40). From the identity and the explicit expression for cos A given immediately above

Since the right hand side is invariant under a cyclic permutation of a, b, and c the spherical sine rule follows immediately.

Alternative derivations

There are many ways of deriving the fundamental cosine and sine rules and the other rules developed in the following sections. For example, Todhunter [1] gives two proofs of the cosine rule (Articles 37 and 60) and two proofs of the sine rule (Articles 40 and 42). The page on Spherical law of cosines gives four different proofs of the cosine rule. Text books on geodesy [2] and spherical astronomy [3] give different proofs and the online resources of MathWorld provide yet more. [4] There are even more exotic derivations, such as that of Banerjee [5] who derives the formulae using the linear algebra of projection matrices and also quotes methods in differential geometry and the group theory of rotations.

The derivation of the cosine rule presented above has the merits of simplicity and directness and the derivation of the sine rule emphasises the fact that no separate proof is required other than the cosine rule. However, the above geometry may be used to give an independent proof of the sine rule. The scalar triple product, OA · (OB × OC) evaluates to sin b sin c sin A in the basis shown. Similarly, in a basis oriented with the z-axis along OB, the triple product OB · (OC × OA), evaluates to sin c sin a sin B. Therefore, the invariance of the triple product under cyclic permutations gives sin b sin A = sin a sin B which is the first of the sine rules. See curved variations of the law of sines to see details of this derivation.

Identities

Supplemental cosine rules

Applying the cosine rules to the polar triangle gives (Todhunter, [1] Art.47), i.e. replacing A by πa, a by πA etc.,

Cotangent four-part formulae

The six parts of a triangle may be written in cyclic order as (aCbAcB). The cotangent, or four-part, formulae relate two sides and two angles forming four consecutive parts around the triangle, for example (aCbA) or BaCb). In such a set there are inner and outer parts: for example in the set (BaCb) the inner angle is C, the inner side is a, the outer angle is B, the outer side is b. The cotangent rule may be written as (Todhunter, [1] Art.44)

and the six possible equations are (with the relevant set shown at right):

To prove the first formula start from the first cosine rule and on the right-hand side substitute for cos c from the third cosine rule:

The result follows on dividing by sin a sin b. Similar techniques with the other two cosine rules give CT3 and CT5. The other three equations follow by applying rules 1, 3 and 5 to the polar triangle.

Half-angle and half-side formulae

With and

Another twelve identities follow by cyclic permutation.

The proof (Todhunter, [1] Art.49) of the first formula starts from the identity using the cosine rule to express A in terms of the sides and replacing the sum of two cosines by a product. (See sum-to-product identities.) The second formula starts from the identity the third is a quotient and the remainder follow by applying the results to the polar triangle.

Delambre analogies

The Delambre analogies (also called Gauss analogies) were published independently by Delambre, Gauss, and Mollweide in 1807–1809. [6]

Another eight identities follow by cyclic permutation.

Proved by expanding the numerators and using the half angle formulae. (Todhunter, [1] Art.54 and Delambre [7] )

Napier's analogies

Another eight identities follow by cyclic permutation.

These identities follow by division of the Delambre formulae. (Todhunter, [1] Art.52)

Taking quotients of these yields the law of tangents, first stated by Persian mathematician Nasir al-Din al-Tusi (1201–1274),

Napier's rules for right spherical triangles

Spherical trigonometry Napier right-angled.svg

When one of the angles, say C, of a spherical triangle is equal to π/2 the various identities given above are considerably simplified. There are ten identities relating three elements chosen from the set a, b, c, A, and B.

Napier [8] provided an elegant mnemonic aid for the ten independent equations: the mnemonic is called Napier's circle or Napier's pentagon (when the circle in the above figure, right, is replaced by a pentagon).

First, write the six parts of the triangle (three vertex angles, three arc angles for the sides) in the order they occur around any circuit of the triangle: for the triangle shown above left, going clockwise starting with a gives aCbAcB. Next replace the parts that are not adjacent to C (that is A, c, and B) by their complements and then delete the angle C from the list. The remaining parts can then be drawn as five ordered, equal slices of a pentagram, or circle, as shown in the above figure (right). For any choice of three contiguous parts, one (the middle part) will be adjacent to two parts and opposite the other two parts. The ten Napier's Rules are given by

The key for remembering which trigonometric function goes with which part is to look at the first vowel of the kind of part: middle parts take the sine, adjacent parts take the tangent, and opposite parts take the cosine. For an example, starting with the sector containing a we have:

The full set of rules for the right spherical triangle is (Todhunter, [1] Art.62)

Napier's rules for quadrantal triangles

A quadrantal spherical triangle together with Napier's circle for use in his mnemonics Spherical trigonometry Napier quadrantal 01.svg
A quadrantal spherical triangle together with Napier's circle for use in his mnemonics

A quadrantal spherical triangle is defined to be a spherical triangle in which one of the sides subtends an angle of π/2 radians at the centre of the sphere: on the unit sphere the side has length π/2. In the case that the side c has length π/2 on the unit sphere the equations governing the remaining sides and angles may be obtained by applying the rules for the right spherical triangle of the previous section to the polar triangle A'B'C' with sides a', b', c' such that A' = πa, a' = πA etc. The results are:

Five-part rules

Substituting the second cosine rule into the first and simplifying gives:

Cancelling the factor of sin c gives

Similar substitutions in the other cosine and supplementary cosine formulae give a large variety of 5-part rules. They are rarely used.

Cagnoli's Equation

Multiplying the first cosine rule by cos A gives

Similarly multiplying the first supplementary cosine rule by cos a yields

Subtracting the two and noting that it follows from the sine rules that produces Cagnoli's equation

which is a relation between the six parts of the spherical triangle. [9]

Solution of triangles

Oblique triangles

The solution of triangles is the principal purpose of spherical trigonometry: given three, four or five elements of the triangle, determine the others. The case of five given elements is trivial, requiring only a single application of the sine rule. For four given elements there is one non-trivial case, which is discussed below. For three given elements there are six cases: three sides, two sides and an included or opposite angle, two angles and an included or opposite side, or three angles. (The last case has no analogue in planar trigonometry.) No single method solves all cases. The figure below shows the seven non-trivial cases: in each case the given sides are marked with a cross-bar and the given angles with an arc. (The given elements are also listed below the triangle). In the summary notation here such as ASA, A refers to a given angle and S refers to a given side, and the sequence of A's and S's in the notation refers to the corresponding sequence in the triangle.

Spherical trigonometry triangle cases.svg

The solution methods listed here are not the only possible choices: many others are possible. In general it is better to choose methods that avoid taking an inverse sine because of the possible ambiguity between an angle and its supplement. The use of half-angle formulae is often advisable because half-angles will be less than π/2 and therefore free from ambiguity. There is a full discussion in Todhunter. The article Solution of triangles#Solving spherical triangles presents variants on these methods with a slightly different notation.

There is a full discussion of the solution of oblique triangles in Todhunter. [1] :Chap. VI See also the discussion in Ross. [10] Nasir al-Din al-Tusi was the first to list the six distinct cases (2-7 in the diagram) of a right triangle in spherical trigonometry. [11]

Spherical trigonometry solution construction.svg

Solution by right-angled triangles

Another approach is to split the triangle into two right-angled triangles. For example, take the Case 3 example where b, c, and B are given. Construct the great circle from A that is normal to the side BC at the point D. Use Napier's rules to solve the triangle ABD: use c and B to find the sides AD and BD and the angle BAD. Then use Napier's rules to solve the triangle ACD: that is use AD and b to find the side DC and the angles C and DAC. The angle A and side a follow by addition.

Numerical considerations

Not all of the rules obtained are numerically robust in extreme examples, for example when an angle approaches zero or π. Problems and solutions may have to be examined carefully, particularly when writing code to solve an arbitrary triangle.

Area and spherical excess

Lexell's theorem: the triangles of constant area on a fixed base AB have their free vertex C along a small circle through the points antipodal to A and B. Lexell's theorem.png
Lexell's theorem: the triangles of constant area on a fixed base AB have their free vertex C along a small circle through the points antipodal to A and B.

Consider an N-sided spherical polygon and let An denote the n-th interior angle. The area of such a polygon is given by (Todhunter, [1] Art.99)

For the case of a spherical triangle with angles A, B, and C this reduces to Girard's theorem

where E is the amount by which the sum of the angles exceeds π radians, called the spherical excess of the triangle. This theorem is named after its author, Albert Girard. [12] An earlier proof was derived, but not published, by the English mathematician Thomas Harriot. On a sphere of radius R both of the above area expressions are multiplied by R2. The definition of the excess is independent of the radius of the sphere.

The converse result may be written as

Since the area of a triangle cannot be negative the spherical excess is always positive. It is not necessarily small, because the sum of the angles may attain 5π (3π for proper angles). For example, an octant of a sphere is a spherical triangle with three right angles, so that the excess is π/2. In practical applications it is often small: for example the triangles of geodetic survey typically have a spherical excess much less than 1' of arc. (Rapp [13] Clarke, [14] Legendre's theorem on spherical triangles). On the Earth the excess of an equilateral triangle with sides 21.3 km (and area 393 km2) is approximately 1 arc second.

There are many formulae for the excess. For example, Todhunter, [1] (Art.101—103) gives ten examples including that of L'Huilier:

where

Because some triangles are badly characterized by their edges (e.g., if ), it is often better to use the formula for the excess in terms of two edges and their included angle

When triangle ABC is a right triangle with right angle at C, then cos C = 0 and sin C = 1, so this reduces to

Angle deficit is defined similarly for hyperbolic geometry.

From latitude and longitude

The spherical excess of a spherical quadrangle bounded by the equator, the two meridians of longitudes and and the great-circle arc between two points with longitude and latitude and is

This result is obtained from one of Napier's analogies. In the limit where are all small, this reduces to the familiar trapezoidal area, .

The area of a polygon can be calculated from individual quadrangles of the above type, from (analogously) individual triangle bounded by a segment of the polygon and two meridians, [15] by a line integral with Green's theorem, [16] or via an equal-area projection as commonly done in GIS. The other algorithms can still be used with the side lengths calculated using a great-circle distance formula.

See also

Related Research Articles

<span class="mw-page-title-main">John Napier</span> Scottish mathematician (1550–1617)

John Napier of Merchiston, nicknamed Marvellous Merchiston, was a Scottish landowner known as a mathematician, physicist, and astronomer. He was the 8th Laird of Merchiston. His Latinized name was Ioannes Neper.

<span class="mw-page-title-main">Trigonometric functions</span> Functions of an angle

In mathematics, the trigonometric functions are real functions which relate an angle of a right-angled triangle to ratios of two side lengths. They are widely used in all sciences that are related to geometry, such as navigation, solid mechanics, celestial mechanics, geodesy, and many others. They are among the simplest periodic functions, and as such are also widely used for studying periodic phenomena through Fourier analysis.

<span class="mw-page-title-main">Law of sines</span> Property of all triangles on a Euclidean plane

In trigonometry, the law of sines, sine law, sine formula, or sine rule is an equation relating the lengths of the sides of any triangle to the sines of its angles. According to the law,

<span class="mw-page-title-main">Law of tangents</span> Relates tangents of two angles of a triangle and the lengths of the opposing sides

In trigonometry, the law of tangents or tangent rule is a statement about the relationship between the tangents of two angles of a triangle and the lengths of the opposing sides.

<span class="mw-page-title-main">Inverse trigonometric functions</span> Inverse functions of sin, cos, tan, etc.

In mathematics, the inverse trigonometric functions are the inverse functions of the trigonometric functions. Specifically, they are the inverses of the sine, cosine, tangent, cotangent, secant, and cosecant functions, and are used to obtain an angle from any of the angle's trigonometric ratios. Inverse trigonometric functions are widely used in engineering, navigation, physics, and geometry.

<span class="mw-page-title-main">Tangent half-angle formula</span> Relates the tangent of half of an angle to trigonometric functions of the entire angle

In trigonometry, tangent half-angle formulas relate the tangent of half of an angle to trigonometric functions of the entire angle. The tangent of half an angle is the stereographic projection of the circle through the point at angle radians onto the line through the angles . Among these formulas are the following:

The Pythagorean trigonometric identity, also called simply the Pythagorean identity, is an identity expressing the Pythagorean theorem in terms of trigonometric functions. Along with the sum-of-angles formulae, it is one of the basic relations between the sine and cosine functions.

<span class="mw-page-title-main">Hyperbolic triangle</span> Triangle in hyperbolic geometry

In hyperbolic geometry, a hyperbolic triangle is a triangle in the hyperbolic plane. It consists of three line segments called sides or edges and three points called angles or vertices.

<span class="mw-page-title-main">Spherical law of cosines</span> Mathematical relation in spherical triangles

In spherical trigonometry, the law of cosines is a theorem relating the sides and angles of spherical triangles, analogous to the ordinary law of cosines from plane trigonometry.

<span class="mw-page-title-main">Great-circle navigation</span> Flight or sailing route along the shortest path between two points on a globes surface

Great-circle navigation or orthodromic navigation is the practice of navigating a vessel along a great circle. Such routes yield the shortest distance between two points on the globe.

<span class="mw-page-title-main">Sine and cosine</span> Fundamental trigonometric functions

In mathematics, sine and cosine are trigonometric functions of an angle. The sine and cosine of an acute angle are defined in the context of a right triangle: for the specified angle, its sine is the ratio of the length of the side that is opposite that angle to the length of the longest side of the triangle, and the cosine is the ratio of the length of the adjacent leg to that of the hypotenuse. For an angle , the sine and cosine functions are denoted simply as and .

<span class="mw-page-title-main">History of trigonometry</span>

Early study of triangles can be traced to the 2nd millennium BC, in Egyptian mathematics and Babylonian mathematics. Trigonometry was also prevalent in Kushite mathematics. Systematic study of trigonometric functions began in Hellenistic mathematics, reaching India as part of Hellenistic astronomy. In Indian astronomy, the study of trigonometric functions flourished in the Gupta period, especially due to Aryabhata, who discovered the sine function, cosine function, and versine function.

There are several equivalent ways for defining trigonometric functions, and the proofs of the trigonometric identities between them depend on the chosen definition. The oldest and most elementary definitions are based on the geometry of right triangles. The proofs given in this article use these definitions, and thus apply to non-negative angles not greater than a right angle. For greater and negative angles, see Trigonometric functions.

<span class="mw-page-title-main">Differentiation of trigonometric functions</span> Mathematical process of finding the derivative of a trigonometric function

The differentiation of trigonometric functions is the mathematical process of finding the derivative of a trigonometric function, or its rate of change with respect to a variable. For example, the derivative of the sine function is written sin′(a) = cos(a), meaning that the rate of change of sin(x) at a particular angle x = a is given by the cosine of that angle.

<span class="mw-page-title-main">Trigonometry</span> Area of geometry, about angles and lengths

Trigonometry is a branch of mathematics concerned with relationships between angles and side lengths of triangles. In particular, the trigonometric functions relate the angles of a right triangle with ratios of its side lengths. The field emerged in the Hellenistic world during the 3rd century BC from applications of geometry to astronomical studies. The Greeks focused on the calculation of chords, while mathematicians in India created the earliest-known tables of values for trigonometric ratios such as sine.

<span class="mw-page-title-main">Law of cosines</span> Property of all triangles on a Euclidean plane

In trigonometry, the law of cosines relates the lengths of the sides of a triangle to the cosine of one of its angles. For a triangle with sides and opposite respective angles and , the law of cosines states:

<span class="mw-page-title-main">Mollweide's formula</span> Relation between sides and angles of a triangle

In trigonometry, Mollweide's formula is a pair of relationships between sides and angles in a triangle.

Solution of triangles is the main trigonometric problem of finding the characteristics of a triangle, when some of these are known. The triangle can be located on a plane or on a sphere. Applications requiring triangle solutions include geodesy, astronomy, construction, and navigation.

In integral calculus, the tangent half-angle substitution is a change of variables used for evaluating integrals, which converts a rational function of trigonometric functions of into an ordinary rational function of by setting . This is the one-dimensional stereographic projection of the unit circle parametrized by angle measure onto the real line. The general transformation formula is:

References

  1. 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 Todhunter, I. (1886). Spherical Trigonometry (5th ed.). MacMillan. Archived from the original on 2020-04-14. Retrieved 2013-07-28.
  2. Clarke, Alexander Ross (1880). Geodesy. Oxford: Clarendon Press. OCLC   2484948 via the Internet Archive.
  3. Smart, W.M. (1977). Text-Book on Spherical Astronomy (6th ed.). Cambridge University Press. Chapter 1 via the Internet Archive.
  4. Weisstein, Eric W. "Spherical Trigonometry". MathWorld . Retrieved 8 April 2018.
  5. Banerjee, Sudipto (2004), "Revisiting Spherical Trigonometry with Orthogonal Projectors", The College Mathematics Journal, 35 (5), Mathematical Association of America: 375–381, doi:10.1080/07468342.2004.11922099, JSTOR   4146847, S2CID   122277398, archived from the original on 2020-07-22, retrieved 2016-01-10
  6. Todhunter, Isaac (1873). "Note on the history of certain formulæ in spherical trigonometry". The London, Edinburgh, and Dublin Philosophical Magazine and Journal of Science. 45 (298): 98–100. doi:10.1080/14786447308640820.
  7. Delambre, J. B. J. (1807). Connaissance des Tems 1809. p. 445. Archived from the original on 2020-07-22. Retrieved 2016-05-14.
  8. Napier, J (1614). Mirifici Logarithmorum Canonis Constructio. p. 50. Archived from the original on 2013-04-30. Retrieved 2016-05-14. An 1889 translation The Construction of the Wonderful Canon of Logarithms is available as en e-book from Abe Books Archived 2020-03-03 at the Wayback Machine
  9. Chauvenet, William (1867). A Treatise on Plane and Spherical Trigonometry. Philadelphia: J. B. Lippincott & Co. p. 165. Archived from the original on 2021-07-11. Retrieved 2021-07-11.
  10. Ross, Debra Anne. Master Math: Trigonometry, Career Press, 2002.
  11. http://www-history.mcs.st-andrews.ac.uk/Biographies/Al-Tusi_Nasir.html,"One[ permanent dead link ] of al-Tusi's most important mathematical contributions was the creation of trigonometry as a mathematical discipline in its own right rather than as just a tool for astronomical applications. In Treatise on the quadrilateral al-Tusi gave the first extant exposition of the whole system of plane and spherical trigonometry. This work is really the first in history on trigonometry as an independent branch of pure mathematics and the first in which all six cases for a right-angled spherical triangle are set forth"/
  12. Another proof of Girard's theorem may be found at Archived 2012-10-31 at the Wayback Machine .
  13. Rapp, Richard H. (1991). Geometric Geodesy Part I (PDF). p. 89.[ permanent dead link ] (pdf page 99),
  14. Clarke, Alexander Ross (1880). Geodesy. Clarendon Press. (Chapters 2 and 9). Recently republished at Forgotten Books Archived 2020-10-03 at the Wayback Machine
  15. Chamberlain, Robert G.; Duquette, William H. (17 April 2007). Some algorithms for polygons on a sphere. Association of American Geographers Annual Meeting. NASA JPL. Archived from the original on 22 July 2020. Retrieved 7 August 2020.
  16. "Surface area of polygon on sphere or ellipsoid – MATLAB areaint". www.mathworks.com. Archived from the original on 2021-05-01. Retrieved 2021-05-01.