Вы находитесь на странице: 1из 793

Reference ellipsoid

From Wikipedia, the free encyclopedia

Flattened sphere

Geodesy

Fundamentals

 Geodesy
 Geodynamics
 Geomatics
 Cartography
 History

Concepts

 Geographical distance
 Geoid
 Figure of the Earth
 Geodetic datum
 Geodesic
 Geographic coordinate system
 Horizontal position representation
 Latitude / Longitude
 Map projection
 Reference ellipsoid
 Satellite geodesy
 Spatial reference system

Technologies

 Global Navigation Satellite System (GNSS)


 Global Positioning System (GPS)
 GLONASS (Russian)
 BeiDou (BDS) (Chinese)
 Galileo (European)
 Indian Regional Navigation
Satellite System (IRNSS)

Standards

ED50 European Datum 1950


SAD69 South American Datum 1969
GRS 80 Geodetic Reference System 1980
NAD83 North American Datum 1983
WGS84 World Geodetic System 1984
NAVD88 N. American Vertical Datum 1988
ETRS89 European Terrestrial Reference
System 1989
GCJ-02 Chinese encrypted datum 2002

 International Terrestrial Reference System


 Spatial Reference System Identifier (SRID)
 Universal Transverse Mercator (UTM)

History

 NGVD29 (Sea Level Datum 1929)

 v
 t
 e
In geodesy, a reference ellipsoid is a mathematically defined surface that approximates the geoid,
the truer figure of the Earth, or other planetary body. Because of their relative simplicity, reference
ellipsoids are used as a preferred surface on which geodetic network computations are performed
and point coordinates such as latitude, longitude, and elevation are defined.

Contents
[hide]

 1Ellipsoid parameters
 2Coordinates
 3Historical Earth ellipsoids
 4Ellipsoids for other planetary bodies
 5See also
 6Notes
 7References
 8External links

Ellipsoid parameters[edit]
In 1687 Isaac Newton published the Principia in which he included a proof[1][not in citation given] that a rotating
self-gravitating fluid body in equilibrium takes the form of an oblate ellipsoid of revolution which he
termed an oblatespheroid. Current practice[2][3] uses the word 'ellipsoid' alone in preference to the full
term 'oblate ellipsoid of revolution' or the older term 'oblate spheroid'. In the rare instances
(some asteroids and planets) where a more general ellipsoid shape is required as a model the term
used is triaxial (or scalene) ellipsoid. A great many ellipsoids have been used with various sizes and
centres but modern (post-GPS) ellipsoids are centred at the actual center of mass of the Earth or
body being modeled.
The shape of an (oblate) ellipsoid (of revolution) is determined by the shape parameters of
that ellipse which generates the ellipsoid when it is rotated about its minor axis. The semi-major
axis of the ellipse, a, is identified as the equatorial radius of the ellipsoid: the semi-minor axis of the
ellipse, b, is identified with the polar distances (from the centre). These two lengths completely
specify the shape of the ellipsoid but in practice geodesy publications classify reference ellipsoids by
giving the semi-major axis and theinverse flattening, 1/f, The flattening, f, is simply a measure of how
much the symmetry axis is compressed relative to the equatorial radius:

f=(a-b)/a

For the Earth, is around 1/300 corresponding to a difference of the major and minor semi-
axes of approximately 21 km. Some precise values are given in the table below and also
in Figure of the Earth. For comparison, Earth'sMoon is even less elliptical, with a flattening of
less than 1/825, while Jupiter is visibly oblate at about 1/15 and one of Saturn's triaxial
moons, Telesto, is nearly 1/3 to 1/2.
A great many other parameters are used in geodesy but they can all be related to one or two of
the set a, b and f. They are listed in ellipse.

Coordinates[edit]
This section does not cite any sources. Please help
improve this section by adding citations to reliable
sources. Unsourced material may be challenged
and removed. (October 2011) (Learn how and when to remove
this template message)
Main articles: Latitude and Longitude
A primary use of reference ellipsoids is to serve as a basis for a coordinate system
of latitude (north/south), longitude(east/west), and elevation (height). For this purpose it is
necessary to identify a zero meridian, which for Earth is usually thePrime Meridian. For other
bodies a fixed surface feature is usually referenced, which for Mars is the meridian passing
through the crater Airy-0. It is possible for many different coordinate systems to be defined upon
the same reference ellipsoid.
The longitude measures the rotational angle between the zero meridian and the measured point.
By convention for the Earth, Moon, and Sun it is expressed in degrees ranging from −180° to
+180° For other bodies a range of 0° to 360° is used.
The latitude measures how close to the poles or equator a point is along a meridian, and is
represented as an angle from −90° to +90°, where 0° is the equator. The common or geodetic
latitude is the angle between the equatorial plane and a line that is normal to the reference
ellipsoid. Depending on the flattening, it may be slightly different from the geocentric
(geographic) latitude, which is the angle between the equatorial plane and a line from the center
of the ellipsoid. For non-Earth bodies the terms planetographic and planetocentric are used
instead.
The coordinates of a geodetic point are customarily stated as geodetic latitude and longitude,
i.e., the direction in space of the geodetic normal containing the point, and the height h of the
point over the reference ellipsoid. See Geodetic system for more detail.

Historical Earth ellipsoids[edit]


Main article: Earth ellipsoid § Historical Earth ellipsoids
Currently the most common reference ellipsoid used, and that used in the context of the Global
Positioning System, is the one defined by WGS 84.
Traditional reference ellipsoids or geodetic datums are defined regionally and therefore non-
geocentric, e.g., ED50. Modern geodetic datums are established with the aid of GPS and will
therefore be geocentric, e.g., WGS 84.

Ellipsoids for other planetary bodies[edit]


Reference ellipsoids are also useful for geodetic mapping of other planetary bodies including
planets, their satellites, asteroids and comet nuclei. Some well observed bodies such as
the Moon and Mars now have quite precise reference ellipsoids.
For rigid-surface nearly-spherical bodies, which includes all the rocky planets and many moons,
ellipsoids are defined in terms of the axis of rotation and the mean surface height excluding any
atmosphere. Mars is actually egg shaped, where its north and south polar radii differ by
approximately 6 km, however this difference is small enough that the average polar radius is
used to define its ellipsoid. The Earth's Moon is effectively spherical, having almost no bulge at
its equator. Where possible a fixed observable surface feature is used when defining a reference
meridian.
For gaseous planets like Jupiter, an effective surface for an ellipsoid is chosen as the equal-
pressure boundary of one bar. Since they have no permanent observable features the choices
of prime meridians are made according to mathematical rules.
Small moons, asteroids, and comet nuclei frequently have irregular shapes. For some of these,
such as Jupiter's Io, a scalene (triaxial) ellipsoid is a better fit than the oblate spheroid. For
highly irregular bodies the concept of a reference ellipsoid may have no useful value, so
sometimes a spherical reference is used instead and points identified by planetocentric latitude
and longitude. Even that can be problematic for non-convex bodies, such as Eros, in that
latitude and longitude don't always uniquely identify a single surface location.

See also[edit]
 Earth ellipsoid
 Earth radius of curvature
 Meridian arc

Notes[edit]
1. Jump up^ Isaac Newton:Principia Book III Proposition XIX Problem III, p. 407 in Andrew Motte
translation, available on line at [1]
2. Jump up^ Torge, W (2001) Geodesy (3rd edition), published by de Gruyter, ISBN 3-11-017072-8
3. Jump up^ Snyder, John P. (1993). Flattening the Earth: Two Thousand Years of Map
Projections. University of Chicago Press. p. 82.ISBN 0-226-76747-7.

References[edit]
 P. K. Seidelmann (Chair), et al. (2005), “Report Of The IAU/IAG Working Group On
Cartographic Coordinates And Rotational Elements: 2003,” Celestial Mechanics and
Dynamical Astronomy, 91, pp. 203–215.
 Web address: http://astrogeology.usgs.gov/Projects/WGCCRE
 OpenGIS Implementation Specification for Geographic information - Simple feature access -
Part 1: Common architecture, Annex B.4. 2005-11-30
 Web address: http://www.opengeospatial.org

External links[edit]
 Geographic coordinate system
 Coordinate systems and transformations (SPENVIS help page)
 Coordinate Systems, Frames and Datums
Categories:
 Geodesy
 Global Positioning System
 Navigation
 Geophysics
 Surveying
Geodesics on an ellipsoid
From Wikipedia, the free encyclopedia

Geodesy

Fundamentals

 Geodesy
 Geodynamics
 Geomatics
 Cartography
 History

Concepts

 Geographical distance
 Geoid
 Figure of the Earth
 Geodetic datum
 Geodesic
 Geographic coordinate system
 Horizontal position representation
 Latitude / Longitude
 Map projection
 Reference ellipsoid
 Satellite geodesy
 Spatial reference system

Technologies

 Global Navigation Satellite System (GNSS)


 Global Positioning System (GPS)
 GLONASS (Russian)
 BeiDou (BDS) (Chinese)
 Galileo (European)
 Indian Regional Navigation
Satellite System (IRNSS)

Standards

ED50 European Datum 1950


SAD69 South American Datum 1969
GRS 80 Geodetic Reference System 1980
NAD83 North American Datum 1983
WGS84 World Geodetic System 1984
NAVD88 N. American Vertical Datum 1988
ETRS89 European Terrestrial Reference
System 1989
GCJ-02 Chinese encrypted datum 2002

 International Terrestrial Reference System


 Spatial Reference System Identifier (SRID)
 Universal Transverse Mercator (UTM)

History

 NGVD29 (Sea Level Datum 1929)

 v
 t
 e

A geodesic on an oblate ellipsoid


The study of geodesics on an ellipsoid arose in connection with geodesy specifically with the
solution of triangulation networks. The figure of the Earth is well approximated by an oblate ellipsoid,
a slightly flattened sphere. Ageodesic is the shortest path between two points on a curved surface,
i.e., the analogue of a straight line on a plane surface. The solution of a triangulation network on an
ellipsoid is therefore a set of exercises in spheroidal trigonometry (Euler 1755).

Isaac Newton

Leonhard Euler
If the Earth is treated as a sphere, the geodesics are great circles (all of which are closed) and the
problems reduce to ones in spherical trigonometry. However, Newton (1687) showed that the effect
of the rotation of the Earth results in its resembling a slightly oblate ellipsoid and, in this case,
the equatorand the meridians are the only closed geodesics. Furthermore, the shortest path
between two points on the equator does not necessarily run along the equator. Finally, if the ellipsoid
is further perturbed to become a triaxial ellipsoid (with three distinct semi-axes), then only three
geodesics are closed and one of these is unstable.
The problems in geodesy are usually reduced to two main cases: the direct problem, given a starting
point and an initial heading, find the position after traveling a certain distance along the geodesic;
and the inverse problem, given two points on the ellipsoid find the connecting geodesic and hence
the shortest distance between them. Because the flattening of the Earth is small, the geodesic
distance between two points on the Earth is well approximated by the great-circle distance using
the mean Earth radius—the relative error is less than 1%. However, the course of the geodesic can
differ dramatically from that of the great circle. As an extreme example, consider two points on the
equator with a longitude difference of 179°59′; while the connecting great circle follows the equator,
the shortest geodesics pass within 180 km of either pole (the flattening makes two symmetric paths
passing close to the poles shorter than the route along the equator).
Aside from their use in geodesy and related fields such as navigation, terrestrial geodesics arise in
the study of the propagation of signals which are confined (approximately) to the surface of the
Earth, for example, sound waves in the ocean (Munk & Forbes 1989) and the radio signals from
lightning (Casper & Bent 1991). Geodesics are used to define some maritime boundaries, which in
turn determine the allocation of valuable resources as such oil and mineral rights. Ellipsoidal
geodesics also arise in other applications; for example, the propagation of radio waves along the
fuselage of an aircraft, which can be roughly modeled as a prolate (elongated) ellipsoid (Kim &
Burnside 1986).
Geodesics are an important intrinsic characteristic of curved surfaces. The sequence of
progressively more complex surfaces, the sphere, an ellipsoid of revolution, and a triaxial ellipsoid,
provide a useful family of surfaces for investigating the general theory of surfaces. Indeed, Gauss's
work on the survey of Hanover, which involved geodesics on an oblate ellipsoid, was a key
motivation for his study of surfaces (Gauss 1828). Similarly, the existence of three closed geodesics
on a triaxial ellipsoid turns out to be a general property of closed, simply connected surfaces;
this theorem of the three geodesics was conjectured by Poincaré (1905) and proved by Lyusternik &
Schnirelmann (1929) (Klingenberg 1982, §3.7).

Contents
[hide]

 1Geodesics on an ellipsoid of revolution


o 1.1Equations for a geodesic
o 1.2Behavior of geodesics
o 1.3Evaluation of the integrals
o 1.4Solution of the direct problem
o 1.5Solution of the inverse problem
o 1.6Differential behavior of geodesics
o 1.7Geodesic map projections
o 1.8Envelope of geodesics
o 1.9Area of a geodesic polygon
o 1.10Software implementations
 2Geodesics on a triaxial ellipsoid
o 2.1Triaxial coordinate systems
o 2.2Jacobi's solution
o 2.3Survey of triaxial geodesics
 3Applications
 4See also
 5Notes
 6References
 7External links

Geodesics on an ellipsoid of revolution[edit]


There are several ways of defining geodesics (Hilbert & Cohn-Vossen 1952, pp. 220–221). A simple
definition is as the shortest path between two points on a surface. However, it is frequently more
useful to define them as paths with zero geodesic curvature—i.e., the analogue of straight lines on a
curved surface. This definition encompasses geodesics traveling so far across the ellipsoid's surface
(somewhat less than half the circumference) that other distinct routes require less distance. Locally,
these geodesics are still identical to the shortest distance between two points.
By the end of the 18th century, an ellipsoid of revolution (the term spheroid is also used) was a well-
accepted approximation to the figure of the Earth. The adjustment of triangulation networks entailed
reducing all the measurements to a reference ellipsoid and solving the resulting two-dimensional
problem as an exercise in spheroidal trigonometry (Bomford 1952, Chap. 3) (Leick et al. 2015, §4.5).
Fig. 1. A geodesic AB on an ellipsoid of revolution. N is the north pole and EFH lie on the equator.
It is possible to reduce the various geodesic problems into one of two types. Consider two
points: A at latitude φ1 and longitude λ1 and B at latitude φ2 and longitude λ2 (see Fig. 1). The
connecting geodesic (from A to B) is AB, of lengths12, which has azimuths α1 and α2 at the two
endpoints.[1] The two geodesic problems usually considered are:

1. the direct geodesic problem or first geodesic problem, given A, α1, and s12,
determine B and α2;
2. the inverse geodesic problem or second geodesic problem, given A and B, determine s12, α1,
and α2.
As can be seen from Fig. 1, these problems involve solving the triangle NAB given one angle, α1 for
the direct problem and λ12 = λ2 − λ1 for the inverse problem, and its two adjacent sides. In the
course of the 18th century these problems were elevated (especially in literature in the German
language) to the principal geodesic problems (Hansen 1865, p. 69).
For a sphere the solutions to these problems are simple exercises in spherical trigonometry, whose
solution is given byformulas for solving a spherical triangle. (See the article on great-circle
navigation.)

Alexis Clairaut
Barnaba Oriani
For an ellipsoid of revolution, the characteristic constant defining the geodesic was found byClairaut
(1735). A systematic solution for the paths of geodesics was given by Legendre (1806)and Oriani
(1806) (and subsequent papers in 1808 and 1810). The full solution for the direct problem (complete
with computational tables and a worked out example) is given by Bessel (1825).[2]
Much of the early work on these problems was carried out by mathematicians—for
example,Legendre, Bessel, and Gauss—who were also heavily involved in the practical aspects
ofsurveying. Beginning in about 1830, the disciplines diverged: those with an interest in geodesy
concentrated on the practical aspects such as approximations suitable for field work, while
mathematicians pursued the solution of geodesics on a triaxial ellipsoid, the analysis of the stability
of closed geodesics, etc.
During the 18th century geodesics were typically referred to as "shortest lines".[3] The term "geodesic
line" was coined by Laplace (1799b):
Nous désignerons cette ligne sous le nom de ligne géodésique [We will call this line the geodesic
line].
This terminology was introduced into English either as "geodesic line" or as "geodetic line", for
example (Hutton 1811),
A line traced in the manner we have now been describing, or deduced from trigonometrical
measures, by the means we have indicated, is called a geodetic orgeodesic line: it has the property
of being the shortest which can be drawn between its two extremities on the surface of the Earth;
and it is therefore the proper itinerary measure of the distance between those two points.
In its adoption by other fields "geodesic line", frequently shortened, to "geodesic", was preferred.[4]
This section treats the problem on an ellipsoid of revolution (both oblate and prolate). The problem
on a triaxial ellipsoid is covered in the next section.
When determining distances on the earth, various approximate methods are frequently used; some
of these are described in the article on geographical distance.
Equations for a geodesic[edit]
Friedrich Bessel

Fig. 2. Differential element of a meridian ellipse.

Fig. 3. Differential element of a geodesic on an ellipsoid.


Here the equations for a geodesic are developed; these allow the geodesics of any length to be
computed accurately. The following derivation closely follows that of Bessel (1825).Bagratuni (1962,
§15), Gan'shin (1967, Chap. 5), Krakiwsky & Thomson (1974, §4), Rapp (1993, §1.2), and Borre &
Strang (2012) also provide derivations of these equations.
Consider an ellipsoid of revolution with equatorial radius a and polar semi-axis b. Define the
flattening f = (a − b)/a, the eccentricity e2 = f(2 − f), and the second eccentricity e′ = e/(1 − f). (In
most applications in geodesy, the ellipsoid is taken to be oblate, a > b; however, the theory applies
without change to prolate ellipsoids, a < b, in which case f, e2, and e′2 are negative.)
Let an elementary segment of a path on the ellipsoid have length ds. From Figs. 2 and 3, we see
that if its azimuth is α, then ds is related todφ and dλ by

(1)
where ρ is the meridional radius of curvature, R = ν cosφ is the radius of the circle of latitude φ,
and ν is the normal radius of curvature. The elementary segment is therefore given by

or

where φ′ = dφ/dλ and the Lagrangian function L depends on φ through ρ(φ) and R(φ).
The length of an arbitrary path between (φ1, λ1) and (φ2, λ2) is given by

where φ is a function of λ satisfying φ(λ1) = φ1 and φ(λ2) = φ2. The shortest path or
geodesic entails finding that functionφ(λ) which minimizes s12. This is an exercise in
the calculus of variations and the minimizing condition is given by theBeltrami
identity,

Fig. 4. The geometric construction for parametric latitude, β. A point P at


latitude φ on the meridian (red) is mapped to a point P′ on a sphere of
radius a (shown as a blue circle) by keeping the radius R constant.

Substituting for L and using Eqs. (1) gives


Clairaut (1735) first found this relation, using a geometrical construction; a
similar derivation is presented by Lyusternik (1964, §10).[5] Differentiating
this relation and manipulating the result gives (Jekeli 2012, Eq. (2.95))

This, together with Eqs. (1), leads to a system of ordinary differential


equations for a geodesic (Jordan & Eggert 1941, §7) (Borre & Strang
2012, Eqs. (11.71) and (11.76))

(2)
We can express R in terms of the parametric latitude, β,[6] using

(see Fig. 4 for the geometrical construction), and Clairaut's


relation then becomes

Fig. 5. Geodesic problem mapped to the auxiliary sphere.

Fig. 6. The elementary geodesic problem on the auxiliary


sphere.
This is the sine rule of spherical trigonometry relating two
sides of the triangle NAB (see Fig. 5), NA = 1⁄2π − β1,
and NB = 1⁄2π − β2 and their opposite angles B = π −
α2 and A = α1.
In order to find the relation for the third sideAB = σ12,
the spherical arc length, and included angle N = ω12,
the spherical longitude, it is useful to consider the
triangle NEP representing a geodesic starting at the
equator; see Fig. 6. In this figure, the variables referred to
the auxiliary sphere are shown with the corresponding
quantities for the ellipsoid shown in parentheses.
Quantities without subscripts refer to the arbitrary
point P; E, the point at which the geodesic crosses the
equator in the northward direction, is used as the origin
for σ, s and ω.

Fig. 7. Differential element of a geodesic on a sphere.

If the side EP is extended by moving P infinitesimally (see


Fig. 7), we obtain

(3)

Combining Eqs. (1) and (3) gives differential equations for s and λ
Up to this point, we have not made use of the specific equations for an ellipsoid, and indeed the
derivation applies to an arbitrary surface of revolution.[7] Bessel now specializes to an ellipsoid in
which R and Z are related by

where Z is the height above the equator (see Fig. 4). Differentiating this and setting dR/dZ =
−sinφ/cosφ gives

eliminating Z from these equations, we obtain

This relation between β and φ can be written as

which is the normal definition of the parametric latitude on an ellipsoid. Furthermore, we have
so that the differential equations for the geodesic become

The last step is to use σ as the independent parameter[8] in both of these differential equations and
thereby to express sand λ as integrals. Applying the sine rule to the vertices E and G in the
spherical triangle EGP in Fig. 6 gives

where α0 is the azimuth at E. Substituting this into the equation for ds/dσ and integrating the result
gives

(4)
where

and the limits on the integral are chosen so that s(σ = 0) = 0. Legendre (1811, p. 180) pointed out
that the equation for s is the same as the equation for the arc on an ellipse with semi-axes b(1
+ e′2 cos2α0)1/2 and b. In order to express the equation for λ in terms of σ, we write

which follows from Eq. (3) and Clairaut's relation. This yields

(5)
and the limits on the integrals are chosen so that λ = λ0 at the equator crossing, σ = 0.
In using these integral relations, we allow σ to increase continuously (not restricting it to a
range [−π, π], for example) as the great circle, resp. geodesic, encircles the auxiliary sphere, resp.
ellipsoid. The quantities ω, λ, and s are likewise allowed to increase without limit. Once the problem
is solved, λ can be reduced to the conventional range.
This completes the solution of the path of a geodesic using the auxiliary sphere. By this device a
great circle can be mapped exactly to a geodesic on an ellipsoid of revolution. However, because
the equations for s and λ in terms of the spherical quantities depend on α0, the mapping is not a
consistent mapping of the surface of the sphere to the ellipsoid or vice versa; instead, it should be
viewed merely as a convenient tool for solving for a particular geodesic.
There are also several ways of approximating geodesics on an ellipsoid which usually apply for
sufficiently short lines (Rapp 1991, §6); however, these are typically comparable in complexity to the
method for the exact solution given above (Jekeli 2012, §2.1.4).
Behavior of geodesics[edit]
Fig. 8. Meridians and the equator are the only closed geodesics. (For the very flattened ellipsoids, there are
other closed geodesics; see Figs. 13 and 14).
Geodesic on an oblate ellipsoid (f = 1⁄50) with α0 = 45°.

Fig. 9. Latitude as a function of longitude for a single cycle of the geodesic from one northward equatorial
crossing to the next.

Fig. 10. Following the geodesic on the ellipsoid for about 5 circuits.
Fig. 11. The same geodesic after about 70 circuits.

Fig. 12. Geodesic on a prolate ellipsoid (f = −1⁄50) with α0 = 45°. Compare with Fig. 10.
Before solving for the geodesics, it is worth reviewing their behavior. Fig. 8 shows the simple
closed geodesics which consist of the meridians (green) and the equator (red). (Here the
qualification "simple" means that the geodesic closes on itself without an intervening self-
intersection.) This follows from the equations for the geodesics given in the previous section.
For meridians, we have α0 = 0 and Eq. (5) becomes λ = ω + λ0, i.e., the longitude will vary the
same way as for a sphere, jumping by π each time the geodesic crosses the pole. The
distance, Eq. (4), reduces to the length of an arc of an ellipse with semi-axes a and b (as
expected), expressed in terms of parametric latitude, β.
The equator (β = 0 on the auxiliary sphere, φ = 0 on the ellipsoid) corresponds toα0 = 1⁄2π. The
distance reduces to the arc of a circle of radius b (and not a), s = bσ, while the longitude
simplifies to λ = (1 − f)σ + λ0. A geodesic that is nearly equatorial will intersect the equator at
intervals of πb. As a consequence, the maximum length of a equatorial geodesic which is also a
shortest path is πb on an oblate ellipsoid (on a prolate ellipsoid, the maximum length is πa).
All other geodesics are typified by Figs. 9 to 11. Figure 9 shows latitude as a function of
longitude for a geodesic starting on the equator with α0 = 45°. A full cycle of the geodesic, from
one northward crossing of the equator to the next, is shown. The equatorial crossings are
called nodes and the points of maximum or minimum latitude are called vertices; the vertex
latitudes are given by|β| = ±(1⁄2π − |α0|). The latitude is an odd, resp. even, function of the
longitude about the nodes, resp. vertices. The geodesic completes one full oscillation in latitude
before the longitude has increased by 360°. Thus, on each successive northward crossing of
the equator (see Fig. 10), λ falls short of a full circuit of the equator by
approximately 2π f sinα0 (for a prolate ellipsoid, this quantity is negative and λ completes more
that a full circuit; see Fig. 12). For nearly all values of α0, the geodesic will fill that portion of the
ellipsoid between the two vertex latitudes (see Fig. 11).
Two additional closed geodesics for the oblate ellipsoid, b/a = 2⁄7.
Fig. 13. Side view.

Fig. 14. Top view.

If the ellipsoid is sufficiently oblate, i.e., b/a < 1⁄2, another class of simple closed geodesics is
possible (Klingenberg 1982, §3.5.19). Two such geodesics are illustrated in Figs. 13 and 14.
Here b/a = 2⁄7 and the equatorial azimuth, α0, for the green (resp. blue) geodesic is chosen to
be 53.175° (resp. 75.192°), so that the geodesic completes 2 (resp. 3) complete oscillations
about the equator on one circuit of the ellipsoid.
Evaluation of the integrals[edit]
Solving the geodesic problems entails evaluating the integrals for the distance, s, and the
longitude, λ, Eqs. (4) and (5). In geodetic applications, where f is small, the integrals are
typically evaluated as a series; for this purpose, the second form of the longitude integral is
preferred (since it avoids the near singular behavior of the first form when geodesics pass
close to a pole). In both integrals, the integrand is an even periodic function of period π.
Furthermore, the term dependent on σ is multiplied by a small quantity k2 = O(f). As a
consequence, the integrals can both be written in the form

where B0 = 1 + O(f) and Bj = O(f j). Series expansions for Bj can readily be found and the
result truncated so that only terms which are O(f J) and larger are retained.[9] (Because the
longitude integral is multiplied by f, it is typically only necessary to retain terms up
to O(f J−1) in that integral.) This prescription is followed by many authors (Legendre 1806)
(Oriani 1806) (Bessel 1825) (Helmert 1880) (Rainsford 1955) (Rapp 1993). Vincenty
(1975a) uses J = 3 which provides an accuracy of about 0.1 mm for
the WGS84 ellipsoid. Karney (2013) gives expansions carried out for J = 6 which suffices to
provide full double precisionaccuracy for |f| ≤ 1⁄50. Trigonometric series of this type can be
conveniently summed using Clenshaw summation.
In order to solve the direct geodesic problem, it is necessary to find σ given s. Since the
integrand in the distance integral is positive, this problem has a unique root, which may be
found using Newton's method, noting that the required derivative is just the integrand of the
distance integral. Oriani (1833) instead uses series reversion so that σ can be found without
iteration; Helmert (1880) gives a similar series.[10] The reverted series converges somewhat
slower that the direct series and, if |f| > 1⁄100, Karney (2013, addenda) supplements the
reverted series with one step of Newton's method to maintain accuracy. Vincenty
(1975a) instead relies on a simpler (but slower) function iteration to solve for σ.
It is also possible to evaluate the integrals (4) and (5) by numerical quadrature (Saito 1970)
(Saito 1979) (Sjöberg & Shirazian 2012) or to apply numerical techniques for the solution of
the ordinary differential equations, Eqs. (2) (Kivioja 1971) (Thomas & Featherstone 2005)
(Panou et al. 2013). Such techniques can be used for arbitrary flattening f. However, if f is
small, e.g., |f| ≤ 1⁄50, they do not offer the speed and accuracy of the series expansions
described above. Furthermore, for arbitrary f, the evaluation of the integrals in terms of elliptic
integrals (see below) also provides a fast and accurate solution. On the other hand, Mathar
(2007) has tackled the more complex problem of geodesics on the surface at a constant
altitude, h, above the ellipsoid by solving the corresponding ordinary differential equations,
Eqs. (2) with [ρ, ν] replaced by [ρ + h, ν + h].

A. M. Legendre
Arthur Cayley
Geodesics on an ellipsoid was an early application of elliptic integrals. In particular,Legendre
(1811) writes the integrals, Eqs. (4) and (5), as

(6)

(7)

Where

and and F(φ, k), E(φ, k), and Π(φ, α2, k), are incomplete elliptic integrals in the notation
of DLMF (2010, §19.2(ii)).[11][12] The first formula for the longitude in Eq. (7) follows directly from the
first form of Eq. (5). The second formula in Eq. (7), due to Cayley (1870), is more convenient for
calculation since the elliptic integral appears in a small term. The equivalence of the two forms
follows from DLMF (2010, Eq. (19.7.8)). Fast algorithms for computing elliptic integrals are given
by Carlson (1995)in terms of symmetric elliptic integrals. Equation (6) is conveniently inverted
using Newton's method. The use of elliptic integrals provides a good method of solving the geodesic
problem for |f| > 1⁄50.[13]
Solution of the direct problem[edit]
The basic strategy for solving the geodesic problems on the ellipsoid is to map the problem onto the
auxiliary sphere by converting φ, λ, and s, to β, ω and σ, to solve the corresponding great-circle
problem on the sphere, and to transfer the results back to the ellipsoid.
In implementing this program, we will frequently need to solve the "elementary" spherical triangle
for NEP in Fig. 6 with Preplaced by either A (subscript 1) or B (subscript 2). For this purpose, we
can apply Napier's rules for quadrantal trianglesto the triangle NEP on the auxiliary sphere which
give

We can also stipulate that cosβ ≥ 0 and cosα0 ≥ 0.[14] Implementing this plan for the
direct problem is straightforward. We are given φ1, α1, and s12. From φ1 we obtain β1 (using
the formula for the parametric latitude). We now solve the triangle problem
with P = A and β1 and α1 given to find α0, σ1, and ω1.[15] Use the distance and longitude
equations, Eqs. (4) and (5), together with the known value of λ1, to find s1 and λ0.
Determine s2 = s1 + s12 and invert the distance equation to findσ2. Solve the triangle problem
with P = B and α0 and σ2 given to find β2, ω2, and α2. Convert β2 to φ2 and
substitute σ2and ω2 into the longitude equation to give λ2.[16]The overall method follows the
procedure for solving the direct problem on a sphere. It is essentially the program laid out
byBessel (1825),[17] Helmert (1880, §5.9), and most subsequent authors.Intermediate
points, way-points, on a geodesic can be found by holding φ1 and α1 fixed and solving the
direct problem for several values of s12. Once the first waypoint is found, only the last portion
of the solution (starting with the determination ofs2) needs to be repeated for each new value
of s12.Solution of the inverse problem[edit]The ease with which the direct problem
can be solved results from the fact that given φ1 and α1, we can immediately findα0, the
parameter in the distance and longitude integrals, Eqs. (4) and (5). In the case of the inverse
problem, we are givenλ12, but we cannot easily relate this to the equivalent spherical
angle ω12 because α0 is unknown. Thus, the solution of the problem requires that α0 be found
iteratively. Before tackling this, it is worth understanding better the behavior of geodesics,
this time, keeping the starting point fixed and varying the azimuth.Geodesics from a single
point (f = 1⁄10, φ1 = −30°)

Fig. 15. Geodesics, geodesic circles, and the cut locus.

Fig. 16. The geodesics shown on aplate carrée projection.


17. λ12 as a function of α1 forφ1 = −30° and φ2 = 20°. Suppose point A in the inverse problem
has φ1 = −30° and λ1 = 0°. Fig. 15 shows geodesics (in blue) emanating A with α1 a multiple
of 15° up to the point at which they cease to be shortest paths. (The flattening has been
increased to 1⁄10 in order to accentuate the ellipsoidal effects.) Also shown (in green) are
curves of constants12, which are the geodesic circles centered A. Gauss (1828) showed that,
on any surface, geodesics and geodesic circle intersect at right angles. The red line is thecut
locus, the locus of points which have multiple (two in this case) shortest geodesics from A.
On a sphere, the cut locus is a point. On an oblate ellipsoid (shown here), it is a segment of
the circle of latitude centered on the point antipodalto A, φ = −φ1. The longitudinal extent of
cut locus is approximatelyλ12 ∈ [π − f π cosφ1, π + f π cosφ1]. If A lies on the equator, φ1 =
0, this relation is exact and as a consequence the equator is only a shortest geodesic if|λ12| ≤
(1 − f)π. For a prolate ellipsoid, the cut locus is a segment of the anti-meridian centered on
the point antipodal to A, λ12 = π, and this means that meridional geodesics stop being
shortest paths before the antipodal point is reached. The solution of the inverse problem
involves determining, for a given point B with latitude φ2 and longitude λ2 which blue and
green curves it lies on; this determinesα1 and s12 respectively. In Fig. 16, the ellipsoid has
been "rolled out" onto a plate carrée projection. Suppose φ2 = 20°, the green line in the
figure. Then as α1 is varied between 0° and 180°, the longitude at which the geodesic
intersects φ = φ2varies between 0° and 180° (see Fig. 17). This behavior holds provided
that|φ2| ≤ |φ1| (otherwise the geodesic does not reach φ2 for some values of α1). Thus, the
inverse problem may be solved by determining the value α1 which results in the given value
of λ12 when the geodesic intersects the circle φ = φ2. This suggests the following strategy for
solving the inverse problem (Karney 2013). Assume that the points A and B satisfy

(8) (There is no loss of generality in this assumption, since the symmetries of the
problem can be used to generate any configuration of points from such configurations.)re
"easy" cases, geodesics which lie on a meridian or the equator. Otherwise…Guess a value
of α1.Solve the so-called hybrid geodesic problem, given φ1, φ2, and α1 find λ12, s12, and α2,
corresponding to the firstintersection of the geodesic with the circle φ = φ2.Compare the
resulting λ12 with the desired value and adjust α1 until the two values agree. This completes
the solution.Each of these steps requires some discussion.1. For an oblate ellipsoid, the
shortest geodesic lies on a meridian if either point lies on a pole or if λ12 = 0 or ±π. The
shortest geodesic follows the equator if φ1 = φ2 = 0 and |λ12| ≤ (1 − f)π. For a prolate
ellipsoid, the meridian is no longer the shortest geodesic if λ12 = ±π and the points are close
to antipodal (this will be discussed in the next section). There is no longitudinal restriction on
equatorial geodesics.2. In most cases a suitable starting value of α1 is found by solving the
spherical inverse problem[14]
(9)

Вам также может понравиться