celestial mechanics

celestial mechanics
the branch of astronomy that deals with the application of the laws of dynamics and Newton's law of gravitation to the motions of heavenly bodies.
[1815-25]

* * *

Branch of astronomy that deals with the mathematical theory of the motions of celestial bodies.

Johannes Kepler's laws of planetary motion (1609–19) and Newton's laws of motion (1687) are fundamental to it. In the 18th century, powerful methods of mathematical analysis were generally successful in accounting for the observed motions of bodies in the solar system. One branch of celestial mechanics deals with the effect of gravitation on rotating bodies, with applications to Earth (see tide) and other objects in space. A modern derivation, called orbital mechanics or flight mechanics, deals with the motions of spacecraft under the influence of gravity, thrust, atmospheric drag, and other forces; it is used to calculate trajectories for ascent into space, achieving orbit, rendezvous, descent, and lunar and interplanetary flights.

* * *

Introduction

      in the broadest sense, the application of classical mechanics to the motion of celestial bodies acted on by any of several types of forces. By far the most important force experienced by these bodies, and much of the time the only important force, is that of their mutual gravitational attraction. But other forces can be important as well, such as atmospheric drag on artificial satellites, the pressure of radiation on dust particles, and even electromagnetic forces on dust particles if they are electrically charged and moving in a magnetic field.

      The term celestial mechanics is sometimes assumed to refer only to the analysis developed for the motion of point mass particles moving under their mutual gravitational attractions, with emphasis on the general orbital motions of solar system bodies. The term astrodynamics is often used to refer to the celestial mechanics of artificial satellite motion. Dynamic astronomy is a much broader term, which, in addition to celestial mechanics and astrodynamics, is usually interpreted to include all aspects of celestial body motion (e.g., rotation, tidal evolution, mass and mass distribution determinations for stars and galaxies, fluid motions in nebulas, and so forth).

Historical background

Early theories
      Celestial mechanics has its beginnings in early astronomy in which the motions of the Sun, the Moon, and the five planets visible to the unaided eye—Mercury, Venus, Mars, Jupiter, and Saturn—were observed and analyzed. The word planet is derived from the Greek word for wanderer, and it was natural for some cultures to elevate these objects moving against the fixed background of the sky to the status of gods; this status survives in some sense today in astrology, where the positions of the planets and Sun are thought to somehow influence the lives of individuals on Earth. The divine status of the planets and their supposed influence on human activities may have been the primary motivation for careful, continued observations of planetary motions and for the development of elaborate schemes for predicting their positions in the future.

      The Greek astronomer Ptolemy (who lived in Alexandria about AD 140) proposed a system of planetary motion in which the Earth was fixed at the centre (geocentric system) and all the planets, the Moon, and the Sun orbited around it. As seen by an observer on the Earth, the planets move across the sky at a variable rate. They even reverse their direction of motion occasionally but resume the dominant direction of motion after a while. To describe this variable motion, Ptolemy assumed that the planets revolved around small circles called epicycles at a uniform rate while the centre of the epicyclic circle orbited the Earth on a large circle called a deferent. Other variations in the motion were accounted for by offsetting the centres of the deferent for each planet from the Earth by a short distance. By choosing the combination of speeds and distances appropriately, Ptolemy was able to predict the motions of the planets with considerable accuracy. His scheme was adopted as absolute dogma and survived more than 1,000 years until the time of Copernicus.

      Nicolaus Copernicus (Copernicus, Nicolaus) assumed that the Earth was just another planet that orbited the Sun along with the other planets. He showed that this heliocentric (heliocentric system) (centred on the Sun) model was consistent with all observations and that it was far simpler than Ptolemy's scheme. His belief that planetary motion had to be a combination of uniform circular motions (Copernican system) forced him to include a series of epicycles to match the motions in the noncircular orbits. The epicycles were like terms in the Fourier series that are used to represent planetary motions today. (A Fourier series is an infinite sum of periodic terms that oscillate between positive and negative values in a smooth way, where the frequency of oscillation changes from term to term. They represent better and better approximations to other functions as more and more terms are kept.) Copernicus also determined the relative scale of his heliocentric solar system, with results that are remarkably close to the modern determination.

      Tycho Brahe (Brahe, Tycho) (1546–1601), who was born three years after Copernicus' death and three years after the publication of the latter's heliocentric model of the solar system, still embraced a geocentric model, but he had only the Sun and the Moon orbiting (Tychonic system) the Earth and all the other planets orbiting the Sun. Although this model is mathematically equivalent to the heliocentric model of Copernicus, it represents an unnecessary complication and is physically incorrect. Tycho's greatest contribution was the more than 20 years of celestial observations he collected; his measurements of the positions of the planets and stars had an unprecedented accuracy of approximately 2 arc minutes. (An arc minute is 1/60 of a degree.)

Kepler's laws of planetary motion
      Tycho's observations were inherited by Johannes Kepler (Kepler, Johannes) (1571–1630), who was employed by Tycho shortly before the latter's death. From these precise positions of the planets at correspondingly accurate times, Kepler empirically determined his famous three laws describing planetary motion: (1) the orbits (orbit) of the planets are ellipses (ellipse) with the Sun at one focus; (2) the radial line from the Sun to the planet sweeps out equal areas in equal times; and (3) the ratio of the squares of the periods of revolution around the Sun of any two planets equal the ratio of the cubes of the semimajor axes of their respective orbital ellipses.

   An ellipse (Figure 1—>) is a plane curve defined such that the sum of the distances from any point G on the ellipse to two fixed points (S and S′ in Figure 1—>) is constant. The two points S and S′ are called foci, and the straight line on which these points lie between the extremes of the ellipse at A and P is referred to as the major axis of the ellipse. Hence, GS + GS′ = AP = 2a in Figure 1—>, where a is the semimajor axis of the ellipse. A focus is separated from the centre C of the ellipse by the fractional part of the semimajor axis given by the product ae, where e < 1 is called the eccentricity. Thus, e = 0 corresponds to a circle. If the Sun is at the focus S of the ellipse, the point P at which the planet is closest to the Sun is called the perihelion, and the most distant point in the orbit A is the aphelion. The term helion refers specifically to the Sun as the primary body about which the planet is orbiting. As the points P and A are also called apses, periapse and apoapse are often used to designate the corresponding points in an orbit about any primary body, although more specific terms, such as perigee and apogee for the Earth, are often used to indicate the primary body. If G is the instantaneous location of a planet in its orbit, the angle f, called the true anomaly, locates this point relative to the perihelion P with the Sun (or focus S) as the origin, or vertex, of the angle. The angle u, called the eccentric anomaly, also locates G relative to P but with the centre of the ellipse as the origin rather than the focus S. An angle called the mean anomaly l (not shown in Figure 1) is also measured from P with S as the origin; it is defined to increase uniformly with time and to equal the true anomaly f at perihelion and aphelion.

 Kepler's second law is also illustrated in Figure 1—>. If the time required for the planet to move from P to F is the same as that to move from D to E, the areas of the two shaded regions will be equal according to the second law. The validity of the second law means a planet must have a higher than average velocity near perihelion and a lower than average velocity near aphelion. The angular velocity (rate of change of the angle f) must vary around the orbit in a similar way. The average angular velocity, called the mean motion, is the rate of change of the mean anomaly l defined above.

      The third law can be used to determine the distance of a planet from the Sun if one knows its orbital period, or vice versa. In particular, if time is measured in years and distance in units of the semimajor axis of the Earth's orbit (i.e., the mean distance of the Earth to the Sun, known as an astronomical unit, or AU), the third law can be written τ2 = a3, where τ is the orbital period.

Newton's laws of motion
      The empirical laws of Kepler describe planetary motion, but Kepler made no attempt to define or constrain the underlying physical processes governing the motion. It was Isaac Newton (Newton, Sir Isaac) who accomplished that feat in the late 17th century. Newton defined momentum as being proportional to velocity with the constant of proportionality being defined as mass. (As described earlier, momentum is a vector quantity in the sense that the direction of motion as well as the magnitude is included in the definition.) Newton then defined force (also a vector quantity) in terms of its effect on moving objects and in the process formulated his three laws of motion: (1) The momentum of an object is constant unless an outside force acts on the object; this means that any object either remains at rest or continues uniform motion in a straight line unless acted on by a force. (2) The time rate of change of the momentum of an object is equal to the force acting on the object. (3) For every action (force) there is an equal and opposite reaction (force). The first law is seen to be a special case of the second law. Galileo, the great Italian contemporary of Kepler who adopted the Copernican point of view and promoted it vigorously, anticipated Newton's first two laws with his experiments in mechanics. But it was Newton who defined them precisely, established the basis of classical mechanics, and set the stage for its application as celestial mechanics to the motions of bodies in space.

      According to the second law, a force must be acting on a planet to cause its path to curve toward the Sun. Newton and others noted that the acceleration of a body in uniform circular motion must be directed toward the centre of the circle; furthermore, if several objects were in circular motion around the same centre at various separations r and their periods of revolution varied as r3/2, as Kepler's third law indicated for the planets, then the acceleration—and thus, by Newton's second law, the force as well—must vary as 1/r2. By assuming this attractive force between point masses, Newton showed that a spherically symmetric mass distribution attracted a second body outside the sphere as if all the spherically distributed mass were contained in a point at the centre of the sphere. Thus the attraction of the planets by the Sun was the same as the gravitational (gravitation) force attracting (gravity) objects to the Earth. Newton further concluded that the force of attraction between two massive bodies was proportional to the inverse square of their separation and to the product of their masses, known as the law of universal gravitation (Newton's law of gravitation). Kepler's laws are derivable from Newton's laws of motion with a central force of gravity varying as 1/r2 from a fixed point, and Newton's law of gravity is derivable from Kepler's laws if one assumes Newton's laws of motion.

 In addition to formulating the laws of motion and of gravity, Newton also showed that a point mass moving about a fixed centre of force, which varies as the inverse square of the distance away from the centre, follows an elliptical path if the initial velocity is not too large, a hyperbolic path for high initial velocities, and a parabolic path for intermediate velocities. In other words, a sequence of orbits in Figure 1—> with the perihelion distance SP fixed but with the velocity at P increasing from orbit to orbit is characterized by a corresponding increase in the orbital eccentricity e from orbit to orbit such that e < 1 for bound elliptical orbits, e = 1 for a parabolic orbit, and e > 1 for a hyperbolic orbit. Many comets have nearly parabolic orbits for their first pass into the inner solar system, whereas spacecraft may have nearly hyperbolic orbits relative to a planet they are flying by while they are close to the planet.

      Throughout history, the motion of the planets in the solar system has served as a laboratory to constrain and guide the development of celestial mechanics in particular and classical mechanics in general. In modern times, increasingly precise observations of celestial bodies have been matched by increasingly precise predictions for future positions—a combination that became a test for Newton's law of gravitation itself. Although the lunar motion (within observational errors) seemed consistent with a gravitational attraction between point masses that decreased exactly as 1/r2, this law of gravitation was ultimately shown to be an approximation of the more complete description of gravity given by the theory of general relativity. Similarly, a discrepancy of roughly 40 arc seconds per century between the observed rate of advance of Mercury's perihelion and that predicted by planetary perturbations with Newtonian gravity is almost precisely accounted for with Einstein's general theory of relativity. That this small discrepancy could be confidently asserted as real was a triumph of quantitative celestial mechanics.

Perturbations and problems of two bodies

The approximate nature of Kepler's laws
      The constraints placed on the force for Kepler's laws to be derivable from Newton's laws were that the force must be directed toward a central fixed point and that the force must decrease as the inverse square of the distance. In actuality, however, the Sun, which serves as the source of the major force, is not fixed but experiences small accelerations because of the planets, in accordance with Newton's second and third laws. Furthermore, the planets attract one another, so that the total force on a planet is not just that due to the Sun; other planets perturb (perturbation) the elliptical motion that would have occurred for a particular planet if that planet had been the only one orbiting an isolated Sun. Kepler's laws therefore are only approximate. The motion of the Sun itself means that, even when the attractions by other planets are neglected, Kepler's third law must be replaced by (M + mi2a3, where mi is one of the planetary masses and M is the Sun's mass. That Kepler's laws are such good approximations to the actual planetary motions results from the fact that all the planetary masses are very small compared to that of the Sun. The perturbations of the elliptic motion are therefore small, and the coefficient M + miM for all the planetary masses mi means that Kepler's third law is very close to being true.

      Newton's second law for a particular mass is a second-order differential equation that must be solved for whatever forces may act on the body if its position as a function of time is to be deduced. The exact solution of this equation, which resulted in a derived trajectory that was an ellipse, parabola, or hyperbola, depended on the assumption that there were only two point particles interacting by the inverse square force. Hence, this “gravitational two-body problem” has an exact solution that reproduces Kepler's laws. If one or more additional bodies also interact with the original pair through their mutual gravitational interactions, no exact solution for the differential equations of motion of any of the bodies involved can be obtained. As was noted above, however, the motion of a planet is almost elliptical, since all masses involved are small compared to the Sun. It is then convenient to treat the motion of a particular planet as slightly perturbed elliptical motion and to determine the changes in the parameters of the ellipse that result from the small forces as time progresses. It is the elaborate developments of various perturbation theories and their applications to approximate the exact motions of celestial bodies that has occupied celestial mechanicians since Newton's time.

Perturbations of elliptical motion
 So far the following orbital parameters, or elements, have been used to describe elliptical motion: the orbital semimajor axis a, the orbital eccentricity e, and, to specify position in the orbit relative to the perihelion, either the true anomaly f, the eccentric anomaly u, or the mean anomaly l. Three more orbital elements are necessary to orient the ellipse in space, since that orientation will change because of the perturbations. The most commonly chosen of these additional parameters are illustrated in Figure 2—>, where the reference plane is chosen arbitrarily to be the plane of the ecliptic, which is the plane of the Earth's orbit defined by the path of the Sun on the sky. (For motion of a near-Earth artificial satellite, the most convenient reference plane would be that of the Earth's Equator.) Angle i is the inclination of the orbital plane to the reference plane. The line of nodes is the intersection of the orbit plane with the reference plane, and the ascending node is that point where the planet travels from below the reference plane (south) to above the reference plane (north). The ascending node is described by its angular position measured from a reference point on the ecliptic plane, such as the vernal equinox; the angle Ω is called the longitude of the ascending node. Angle ω (called the argument of perihelion) is the angular distance from the ascending node to the perihelion measured in the orbit plane.

      For the two-body problem, all the orbital parameters a, e, i, Ω, and ω are constants. A sixth constant T, the time of perihelion passage (i.e., any date at which the object in orbit was known to be at perihelion), may be used to replace f, u, or l, and the position of the planet in its fixed elliptic orbit can be determined uniquely at subsequent times. These six constants are determined uniquely by the six initial conditions of three components of the position vector and three components of the velocity vector relative to a coordinate system that is fixed with respect to the reference plane. When small perturbations are taken into account, it is convenient to consider the orbit as an instantaneous ellipse whose parameters are defined by the instantaneous values of the position and velocity vectors, since for small perturbations the orbit is approximately an ellipse. In fact, however, perturbations cause the six formerly constant parameters to vary slowly, and the instantaneous perturbed orbit is called an osculating ellipse; that is, the osculating ellipse is that elliptical orbit that would be assumed by the body if all the perturbing forces were suddenly turned off.

      First-order differential equations describing the variation of the six orbital parameters can be constructed for each planet or other celestial body from the second-order differential equations that result by equating the mass times the acceleration of a body to the sum of all the forces acting on the body (Newton's second law). These equations are sometimes called the Lagrange planetary equations after their derivation by the great Italian-French mathematician Joseph-Louis Lagrange (Lagrange, Joseph-Louis, comte de l'Empire) (1736–1813). As long as the forces are conservative and do not depend on the velocities—i.e., there is no loss of mechanical energy through such processes as friction—they can be derived from partial derivatives of a function of the spatial coordinates only, called the potential energy, whose magnitude depends on the relative separations of the masses.

      In the case where all the forces are derivable from such potential energy, the total energy of a system of any number of particles—i.e, the kinetic energy plus the potential energy—is constant. The kinetic energy of a single particle is one-half its mass times the square of its velocity, and the total kinetic energy is the sum of such expressions for all the particles being considered. The conservation of energy principle is thus expressed by an equation relating the velocities of all the masses to their positions at any time. The partial derivatives of the potential energy with respect to spatial coordinates are transformed into particle derivatives of a disturbing function with respect to the orbital elements in the Lagrange equations, where the disturbing function vanishes if all bodies perturbing the elliptic motion are removed. Like Newton's equations of motion, Lagrange's differential equations are exact, but they can be solved only numerically on a computer or analytically by successive approximations. In the latter process, the disturbing function is represented by a Fourier series, with convergence of the series (successive decrease in size and importance of the terms) depending on the size of the orbital eccentricities and inclinations. Clever changes of variables and other mathematical tricks are used to increase the time span over which the solutions (also represented by series) are good approximations to the real motion. These series solutions usually diverge, but they still represent the actual motions remarkably well for limited periods of time. One of the major triumphs of celestial mechanics using these perturbation techniques was the discovery of Neptune in 1846 from its perturbations of the motion of Uranus.

Examples of perturbations
 Some of the variations in the orbital parameters caused by perturbations can be understood in simple terms. The lunar (Moon) orbit is inclined to the ecliptic plane by about 5°, and the longitude of its ascending node on the ecliptic plane (Ω in Figure 2—>) is observed to regress (Ω decreasing) a complete revolution in 18.61 years. The Sun is the dominant cause of this regression of the lunar node. When the Moon is closer to the Sun than the Earth, the Sun accelerates the Moon slightly more than it accelerates the Earth. This difference in the accelerations is what perturbs the lunar motion around the Earth. The Moon does not fly off in this situation since the acceleration of the Moon toward the Earth is much larger than the difference between the Sun's accelerations of the Earth and the Moon.

 The Sun, of course, is always in the ecliptic plane, since its apparent path among the stars defines the plane. This means that the perturbing acceleration just defined will always be pointed slightly toward the ecliptic plane whenever the Moon is below or above this plane in its orbital motion about the Earth. This tendency to pull the Moon toward the ecliptic plane means that the Moon will cross the plane on each half orbit at a longitude that is slightly smaller than the longitude at which it would have crossed if the Sun had not been there. Thus, the line of nodes will have regressed. The instantaneous rate at which the node regresses varies as the geometry changes during the Moon's motion around the Earth, and during the Earth-Moon system's motion around the Sun, but there is always a net regression. Such a change that is always in the same direction as time increases is called a secular perturbation. Superposed on the secular perturbation of the longitude of the node are periodic perturbations (periodically changing their direction), which are revealed by the fact that the rate of secular regression of the node is not constant in time. The Sun causes a secular increase in the longitude of the lunar perigee (Ω + ω in Figure 2—>) of one complete revolution in 8.85 years, as well as periodic perturbations in the inclination, eccentricity, and mean motion.

      For near-Earth (Earth satellite) artificial satellites, the deviation of the Earth's mass distribution from spherical symmetry is the dominant cause of the perturbations from pure elliptic motion. The most important deviation is the equatorial bulge of the Earth due to its rotation. If, for example, the Earth were a sphere with a ring of mass around its Equator, the ring would give to a satellite whose orbit is inclined to the Equator a component of acceleration toward the Equator plane whenever the satellite was above or below this plane. By an argument similar to that for the Moon acted on by the Sun, this acceleration would cause the line of nodes of a close satellite orbit to regress a little more than 5° per day.

      As a final example, the distribution of continents and oceans and the varying mass densities in the Earth's mantle (the layer underlying the crust) lead to a slight deviation of the Earth's gravitational force field from axial symmetry. Usually this causes only short-period perturbations of low amplitude for near-Earth satellites. However, communications or weather satellites that are meant to maintain a fixed longitude over the Equator (i.e., geostationary satellites, which orbit synchronously with the Earth's rotation) are destabilized by this deviation except at two longitudes. If the axial asymmetry is represented by a slightly elliptical Equator, the difference between the major and minor axis of the ellipse is about 64 metres, with the major axis located about 35° W. A satellite at a position slightly ahead of the long axis of the elliptical Equator will experience a component of acceleration opposite its direction of orbital motion (as if a large mountain were pulling it back). This acceleration makes the satellite fall closer to the Earth and increases its mean motion, causing it to drift further ahead of the axial bulge on the Equator. If the satellite is slightly behind the axial bulge, it experiences an acceleration in the direction of its motion. This makes the satellite move away from the Earth with a decrease in its mean motion, so that it will drift further behind the axial bulge. The synchronous Earth satellites are thus repelled from the long axis of the equatorial ellipse and attracted to the short axis, and compensating accelerations, usually from onboard jets, are required to stabilize a satellite at any longitude other than the two corresponding to the ends of the short axis of the axial bulge. (The jets are actually required for any longitude, as they must also compensate for other perturbations such as radiation pressure.)

The three-body problem (three-body problem)
      The inclusion of solar perturbations of the motion of the Moon results in a “three-body problem” (Earth-Moon-Sun), which is the simplest complication of the completely solvable two-body problem discussed above. When the Earth, Moon, and Sun are considered to be point masses, this particular three-body problem is called “the main problem of the lunar theory,” which has been studied extensively with a variety of methods beginning with Newton. Although the three-body problem has no complete analytic solution in closed form, various series solutions by successive approximations achieve such accuracy that complete theories of the lunar motion must include the effects of the nonspherical mass distributions of both the Earth and the Moon as well as the effects of the planets if the precision of the predicted positions is to approach that of the observations. Most of the schemes for the main problem are partially numerical and therefore apply only to the lunar motion. An exception is the completely analytic work of the French astronomer Charles-Eugène Delaunay (Delaunay, Charles-Eugène) (1816–72), who exploited and developed the most elegant techniques of classical mechanics pioneered by his contemporary, the Irish astronomer and mathematician William R. Hamilton (1805–65). Delaunay could predict the position of the Moon to within its own diameter over a 20-year time span. Since his development was entirely analytic, the work was applicable to the motions of satellites about other planets where the series expansions converged much more rapidly than they did for the application to the lunar motion.

      Delaunay's work on the lunar theory demonstrates some of the influence that celestial mechanics has had on the development of the techniques of classical mechanics. This close link between the development of classical mechanics and its application to celestial mechanics was probably no better demonstrated than in the work of the French mathematician Henri Poincaré (Poincaré, Henri) (1854–1912). Poincaré, along with other great mathematicians such as George D. Birkhoff (1884–1944), Aurel Wintner (1903–58), and Andrey N. Kolmogorov (1903–87), placed celestial mechanics on a more sound mathematical basis and was less concerned with quantitatively accurate prediction of celestial body motion. Poincaré demonstrated that the series solutions in use in celestial mechanics for so long generally did not converge but that they could give accurate descriptions of the motion for significant periods of time in truncated form. The elaborate theoretical developments in celestial and classical mechanics have received more attention recently with the realization that a large class of motions are of an irregular or chaotic nature and require fundamentally different approaches for their description.

The restricted three-body problem
      The simplest form of the three-body problem is called the restricted three-body problem, in which a particle of infinitesimal mass moves in the gravitational field of two massive bodies orbiting according to the exact solution of the two-body problem. (The particle with infinitesimal mass, sometimes called a massless particle, does not perturb the motions of the two massive bodies.) There is an enormous literature devoted to this problem, including both analytic and numerical developments. The analytic work was devoted mostly to the circular, planar restricted three-body problem, where all particles are confined to a plane and the two finite masses are in circular orbits around their centre of mass (a point on the line between the two masses that is closer to the more massive). Numerical developments allowed consideration of the more general problem.

      In the circular problem, the two finite masses are fixed in a coordinate system rotating at the orbital angular velocity, with the origin (axis of rotation) at the centre of mass of the two bodies. Lagrange showed that in this rotating frame there were five stationary points at which the massless particle would remain fixed if placed there. There are three such points lying on the line connecting the two finite masses: one between the masses and one outside each of the masses. The other two stationary points, called the triangular points, are located equidistant from the two finite masses at a distance equal to the finite mass separation. The two masses and the triangular stationary points are thus located at the vertices of equilateral triangles in the plane of the circular orbit.

      There is a constant of the motion in the rotating frame that leads to an equation relating the velocity of the massless particle in this frame to its position. For given values of this constant it is possible to construct curves in the plane on which the velocity vanishes. If such a zero-velocity curve is closed, the particle cannot escape from the interior of the closed zero-velocity curve if placed there with the constant of the motion equal to the value used to construct the curve. These zero-velocity curves can be used to show that the three collinear stationary points are all unstable in the sense that, if the particle is placed at one of these points, the slightest perturbation will cause it to move far away. The triangular points are stable if the ratio of the finite masses is less than 0.04, and the particle would execute small oscillations around one of the triangular points if it were pushed slightly away. Since the mass ratio of Jupiter to the Sun is about 0.001, the stability criterion is satisfied, and Lagrange predicted the presence of the Trojan asteroids (Trojan planets) at the triangular points of the Sun-Jupiter system 134 years before they were observed. Of course, the stability of the triangular points must also depend on the perturbations by any other bodies. Such perturbations are sufficiently small not to destabilize the Trojan asteroids. Single Trojan-like bodies have also been found orbiting at leading and trailing triangular points in the orbit of Saturn's satellite Tethys, at the leading triangular point in the orbit of another Saturnian satellite, Dione, and at the trailing point in the orbit of Mars.

Orbital resonances
      There are stable configurations in the restricted three-body problem that are not stationary in the rotating frame. If, for example, Jupiter and the Sun are the two massive bodies, these stable configurations occur when the mean motions of Jupiter and the small particle—here an asteroid—are near a ratio of small integers. The orbital mean motions are then said to be nearly commensurate, and an asteroid that is trapped near such a mean motion commensurability is said to be in an orbital resonance with Jupiter. For example, the Trojan asteroids librate (oscillate) around the 1:1 orbital resonance (i.e., the orbital period of Jupiter is in a 1:1 ratio with the orbital period of the Trojan asteroids); the asteroid Thule librates around the 4:3 orbital resonance; and several asteroids in the Hilda group librate around the 3:2 orbital resonance. There are several such stable orbital resonances among the satellites of the major planets and one involving Pluto and the planet Neptune. The analysis based on the restricted three-body problem cannot be used for the satellite resonances, however, except for the 4:3 resonance between Saturn's satellites Titan and Hyperion, since the participants in the satellite resonances usually have comparable masses.

      Although the asteroid Griqua librates around the 2:1 resonance with Jupiter, and Alinda librates around the 3:1 resonance, the orbital commensurabilities 2:1, 7:3, 5:2, and 3:1 are characterized by an absence of asteroids in an otherwise rather highly populated, uniform distribution spanning all of the commensurabilities. These are the Kirkwood gaps in the distribution of asteroids, and the recent understanding of their creation and maintenance has introduced into celestial mechanics an entirely new concept of irregular, or chaotic, orbits in a system whose equations of motion are entirely deterministic.

Chaotic orbits
      The French astronomer Michel Hénon and the American astronomer Carl Heiles discovered that when a system exhibiting periodic motion, such as a pendulum, is perturbed by an external force that is also periodic, some initial conditions lead to motions where the state of the system becomes essentially unpredictable (within some range of system states) at some time in the future, whereas initial conditions within some other set produce quasiperiodic or predictable behaviour. The unpredictable behaviour is called chaotic, and initial conditions that produce it are said to lie in a chaotic zone. If the chaotic zone is bounded, in the sense that only limited ranges of initial values of the variables describing the motion lead to chaotic behaviour, the uncertainty in the state of the system in the future is limited by the extent of the chaotic zone; that is, values of the variables in the distant future are completely uncertain only within those ranges of values within the chaotic zone. This complete uncertainty within the zone means the system will eventually come arbitrarily close to any set of values of the variables within the zone if given sufficient time. Chaotic orbits were first realized in the asteroid belt.

      A periodic term in the expansion of the disturbing function for a typical asteroid orbit becomes more important in influencing the motion of the asteroid if the frequency with which it changes sign is very small and its coefficient is relatively large. For asteroids orbiting near a mean motion commensurability with Jupiter, there are generally several terms in the disturbing function with large coefficients and small frequencies that are close but not identical. These “resonant” terms often dominate the perturbations of the asteroid motion so much that all the higher-frequency terms can be neglected in determining a first approximation to the perturbed motion. This neglect is equivalent to averaging the higher-frequency terms to zero; the low-frequency terms change only slightly during the averaging. If one of the frequencies vanishes on the average, the periodic term becomes nearly constant, or secular, and the asteroid is locked into an exact orbital resonance near the particular mean motion commensurability. The mean motions are not exactly commensurate in such a resonance, however, since the motion of the asteroid orbital node or perihelion is always involved (except for the 1:1 Trojan resonances).

 For example, for the 3:1 commensurability, the angle θ = λA - 3λJ + ϖA is the argument of one of the important periodic terms whose variation can vanish (zero frequency). Here λ = Ω + ω + l is the mean longitude, the subscripts A and J refer to the asteroid and Jupiter, respectively, and ϖ = Ω + ω is the longitude of perihelion (see Figure 2—>). Within resonance, the angle θ librates, or oscillates, around a constant value as would a pendulum around its equilibrium position at the bottom of its swing. The larger the amplitude of the equivalent pendulum, the larger its velocity at the bottom of its swing. If the velocity of the pendulum at the bottom of its swing, or, equivalently, the maximum rate of change of the angle θ, is sufficiently high, the pendulum will swing over the top of its support and be in a state of rotation instead of libration. The maximum value of the rate of change of θ for which θ remains an angle of libration (periodically reversing its variation) instead of one of rotation (increasing or decreasing monotonically) is defined as the half-width of the resonance.

      Another term with nearly zero frequency when the asteroid is near the 3:1 commensurability has the argument θ′ = λA - λJ + 2ϖJ. The substitution of the longitude of Jupiter's perihelion for that of the asteroid means that the rates of change of θ and θ′ will be slightly different. As the resonances are not separated much in frequency, there may exist values of the mean motion of the asteroid where both θ and θ′ would be angles of libration if either resonance existed in the absence of the other. The resonances are said to overlap in this case, and the attempt by the system to librate simultaneously about both resonances for some initial conditions leads to chaotic orbital behaviour. The important characteristic of the chaotic zone for asteroid motion near a mean motion commensurability with Jupiter is that it includes a region where the asteroid's orbital eccentricity is large. During the variation of the elements over the entire chaotic zone as time increases, large eccentricities must occasionally be reached. For asteroids near the 3:1 commensurability with Jupiter, the orbit then crosses that of Mars, whose gravitational interaction in a close encounter can remove the asteroid from the 3:1 zone.

      By numerically integrating many orbits whose initial conditions spanned the 3:1 Kirkwood gap region in the asteroid belt, Jack Wisdom, an American dynamicist who developed a powerful means of analyzing chaotic motions, found that the chaotic zone around this gap precisely matched the physical extent of the gap. There are no observable asteroids with orbits within the chaotic zone, but there are many just outside extremes of the zone. Preliminary work has indicated that the other Kirkwood gaps can be similarly accounted for. The realization that orbits governed by Newton's laws of motion and gravitation could have chaotic properties and that such properties could solve a long-standing problem in the celestial mechanics of the solar system is a major breakthrough in the subject.

-body problem
      The general problem of n bodies, where n is greater than three, has been attacked vigorously with numerical techniques on powerful computers. Celestial mechanics in the solar system is ultimately an n-body problem, but the special configurations and relative smallness of the perturbations have allowed quite accurate descriptions of motions (valid for limited time periods) with various approximations and procedures without any attempt to solve the complete problem of n bodies. Examples are the restricted three-body problem to determine the effect of Jupiter's perturbations of the asteroids and the use of successive approximations of series solutions to sequentially add the effects of smaller and smaller perturbations for the motion of the Moon. In the general n-body problem, all bodies have arbitrary masses, initial velocities, and positions; the bodies interact through Newton's law of gravitation, and one attempts to determine the subsequent motion of all the bodies. Many numerical solutions for the motion of quite large numbers of gravitating particles have been successfully completed where the precise motion of individual particles is usually less important than the statistical behaviour of the group.

Numerical (numerical analysis) solutions
      Numerical solutions of the exact equations of motion for n bodies can be formulated. Each body is subject to the gravitational attraction of all the others, and it may be subject to other forces as well. It is relatively easy to write the expression for the instantaneous acceleration (equation of motion) of each body if the position of all the other bodies is known, and expressions for all the other forces can be written (as they can for gravitational forces) in terms of the relative positions of the particles and other defining characteristics of the particle and its environment. Each particle is allowed to move under its instantaneous acceleration for a short time step. Its velocity and position are thereby changed, and the new values of the variables are used to calculate the acceleration for the next time step, and so forth. Of course, the real position and velocity of the particle after each time step will differ from the calculated values by errors of two types. One type results from the fact that the acceleration is not really constant over the time step, and the other from the rounding off or truncation of the numbers at every step of the calculation. The first type of error is decreased by taking shorter time steps. But this means more numerical operations must be carried out over a given span of time, and this increases the round-off error for a given precision of the numbers being carried in the calculation. The design of numerical algorithms, as well as the choice of precisions and step sizes that maximize the speed of the calculation while keeping the errors within reasonable bounds, is almost an art form developed by extensive experience and ingenuity. For example, a scheme exists for extrapolating the step size to zero in order to find the change in the variables over a relatively short time span, thereby minimizing the accumulation of error from this source. If the total energy of the system is theoretically conserved, its evaluation for values of the variables at the beginning and end of a calculation is a measure of the errors that have accumulated.

      The motion of the planets of the solar system over time scales approaching its 4.6-billion-year age is a classic n-body problem, where n = 10 with the Sun included. The question of whether or not the solar system is ultimately stable—whether the current configuration of the planets will be maintained indefinitely under their mutual perturbations, or whether one or another planet will eventually be lost from the system or otherwise have its orbit drastically altered—is a long-standing one that might someday be answered through numerical calculation. The interplay of orbital resonances and chaotic orbits discussed above can be investigated numerically, and this interplay may be crucial in determining the stability of the solar system. Already it appears that the parameters defining the orbits of several planets vary over narrow chaotic zones, but whether or not this chaos can lead to instability if given enough time is still uncertain.

      If accelerations are determined by summing all the pairwise interactions for the n particles, the computer time per time step increases as n2. Practical computations for the direct calculation of the interactions between all the particles are thereby limited to n < 10,000. Therefore, for larger values of n, schemes are used where a particle is assumed to move in the force field of the remaining particles approximated by that due to a continuum mass distribution, or a “tree structure” is used where the effects of nearby particles are considered individually while larger and larger groups of particles are considered collectively as their distances increase. These later schemes have the capability of calculating the evolution of a very large system of particles using a reasonable amount of computer time with reasonable approximation. Values of n near 100,000 have been used in calculations determining the evolution of galaxies of stars. Also, the consequences for distribution of stars when two galaxies closely approach one another or even collide has been determined. Even calculations of the n-body problem where n changes with time have been completed in the study of the accumulation of larger bodies from smaller bodies via collisions in the process of the formation of the planets.

      In all n-body calculations, very close approaches of two particles can result in accelerations so large and so rapidly changing that large errors are introduced or the calculation completely diverges. Accuracy can sometimes be maintained in such a close approach, but only at the expense of requiring very short time steps, which drastically slows the calculation. When n is small, as in some solar system calculations where two-body orbits still dominate, close approaches are sometimes handled by a change to a set of variables, usually involving the eccentric anomaly u, that vary much less rapidly during the encounter. In this process, called regularization, the encounter is traversed in less computer time while preserving reasonable accuracy. This process is impractical when n is large, so accelerations are usually artificially bounded on close approaches to prevent instabilities in the numerical calculation and to prevent slowing the calculation. For example, if several sets of particles were trapped in stable, close binary orbits, the very short time steps required to follow this rapid motion would bring the calculation to a virtual standstill, and such binary motion is not important in the overall evolution of, say, a galaxy of stars.

Algebraic maps
      In numerical calculations for conservative systems with modest values of n over long time spans, such as those seeking a determination of the stability of the solar system, the direct solution of the differential equations governing the motions requires excessive time on any computer and accumulates excessive round-off error in the process. Excessive time also is required to explore thoroughly a complete range of orbital parameters in numerical experiments in order to determine the extent of chaotic zones in various configurations (e.g., those in the asteroid belt near orbital mean motion commensurabilities with Jupiter). A solution to this problem is the use of an algebraic map (mapping), which maps the space of system variables onto itself in such a way that the values of all the variables at one instant of time are expressed by algebraic relations in terms of the values of the variables at a fixed time in the past. The values at the next time step are determined by applying the same map to the values just obtained, and so on. The map is constructed by assuming that the motions of all the bodies are unperturbed for a given short time but are periodically “kicked” by the perturbing forces for only an instant. The continuous perturbations are thus replaced by periodic impulses. The values of the variables are “mapped” from one time step to the next by the fact that the unperturbed part of the motion is available from the exact solution of the two-body problem, and it is easy to solve the equations with all the perturbations over the short time of the impulse. Although this approximation does not produce exactly the same values of all the variables at some time in the future as those produced by a numerical solution of the differential equations starting with the same initial conditions, the qualitative behaviour is indistinguishable over long time periods. As computers can perform the algebraic calculations as much as 1,000 times faster than they can solve the corresponding differential equations, the computational time savings are enormous and problems otherwise impossible to explore become tractable.

Tidal evolution
      This discussion has so far treated the celestial mechanics of bodies accelerated by conservative forces (total energy being conserved), including perturbations of elliptic motion by nonspherical mass distributions of finite-size bodies. However, the gravitational field of one body in close orbit about another will tidally distort (tidal friction) the shape of the other body. Dissipation of part of the energy stored in these tidal distortions leads to a coupling that causes secular changes (always in the same direction) in the orbit and in the spins of both bodies. Since tidal dissipation accounts for the current spin states of several planets, the spin states of most of the planetary satellites and some of their orbital configurations, and the spins and orbits of close binary stars, it is appropriate that tides (tide) and their consequences be included in this discussion.

      The twice-daily high and low tides in the ocean are known by all who have lived near a coast. Few are aware, however, that the solid body of the Earth also experiences twice-daily tides with a maximum amplitude of about 30 centimetres. George Howard Darwin (Darwin, Sir George) (1845–1912), the second son of Charles Darwin, the naturalist, was an astronomer-geophysicist who understood quantitatively the generation of the tides in the gravitational fields of tide-raising bodies, which are primarily the Moon and Sun for the Earth; he pointed out that the dissipation of tidal energy resulted in a slowing of the Earth's rotation while the Moon's orbit was gradually expanded. That any mass raises a tide on every other mass within its gravitational field follows from the fact that the gravitational force between two masses decreases as the inverse square of the distance between them.

  In Figure 3—>, the accelerations due to mass ms of three mass elements in the spherical mass mp are proportional to the length of the arrows attached to each element. The element nearest ms is accelerated (acceleration) more than the element at the centre of mp and tends to leave the centre element behind; the element at the centre of mp is accelerated more than the element farthest from ms, and the latter tends to be left further behind. The point of view of a fictitious observer at the centre of mp can be realized by subtracting the acceleration of the central mass element from that of each of the other two mass elements. If the mass elements were free, this observer would see the two extreme mass elements being accelerated in opposite directions away from his position at the centre, as illustrated in Figure 3B—>.

  But the mass elements are not free; they are gravitationally attracted to one another and to the remaining mass in mp. The gravitational acceleration of the mass elements on the surface of mp toward the centre of mp far exceeds the differential acceleration due to the gravitational attraction of ms, thus the elements do not fly off. If mp were incompressible and perfectly rigid, the mass elements on the surface would weigh a little less than they would if ms were not there but would not move relative to the centre of mp. If mp were fluid or otherwise not rigid, it would distort into an oval shape in the presence of ms. The reason for this distortion is that the mass elements making up mp that do not lie on the line joining the centres of mp and ms also experience a differential acceleration. Such differential accelerations are not perpendicular to the surface, however, and are therefore not compensated by the self-gravity that accelerates mass elements toward the centre of mp. This is shown in Figure 4A—>, where one of the differential accelerations is resolved into two components (dotted arrows), one perpendicular and one tangential to the surface. The perpendicular component is compensated by the self-gravity; the tangential component is not. If mp were entirely fluid, the uncompensated tangential components of the differential accelerations due to ms would cause mass to flow toward the points on mp that were either closest to ms or farthest from ms until mp would resemble Figure 4B—>. In this shape the self-gravity is no longer perpendicular to the surface but has a component opposite the tangential component of the differential acceleration. Only in this distorted shape will all the differential accelerations be compensated and the entire body accelerated like the centre. If mp is not fluid but is rigid like rock or iron, part of the compensating acceleration will be provided by internal stress forces, and the body will distort less. As no material is perfectly rigid, there is always some tidal bulge, and compressibility of the material will farther enhance this bulge. Note that the tidal distortion is independent of the orbital motion and would also occur if mp and ms were simply falling toward each other. (There is a similar tide raised on ms by mp that will be ignored for the present.)

  If mp rotates relative to ms, an observer on the surface of mp would successively rotate through the maxima and minima of the tidal distortion, which would tend to remain aligned with the direction to ms. The observer would thereby experience two high (ebb tide) and two low tides a day, as observed on Earth. Some of the energy of motion of any fluid parts of mp and some of the energy stored as distortion of the solid parts as the tides wax and wane is converted into heat, and this dissipation of mechanical energy causes a delay in the response of the body to the tide-raising force. This means that high tide would occur at a given point on mp as it rotates relative to ms after ms passes overhead. (On the Earth, the continents alter the motion of the fluid ocean so much that ocean tides at continental coasts do not always lag the passage of the Moon overhead.) If mp rotates in the same direction as ms revolves in its orbit, the tidal bulge is carried ahead of ms, as shown in Figure 5—> by angle δ. Again, because the gravitational force between two masses varies as the inverse square of their separation, the tidal bulge closest to ms experiences a greater attraction toward ms (F1 in Figure 5—>) than does the bulge farthest away (F2). As these two forces are not aligned with the centre of mp, there is a twisting effect, or torque, on mp that retards its rate of rotation. This retardation will continue until the rotation is synchronous with the mean orbital motion of ms. This has happened for the Moon, which keeps the same face toward the Earth.

 From Newton's third law, there are equal and opposite forces acting on ms corresponding to F1 and F2. In Figure 5—> these forces are represented as T1 and T2, and each has been resolved into two components, one directed toward the centre of mp and the other perpendicular to this direction. The inequality of these forces causes a net acceleration of ms in its orbit, which thereby expands, as is observed for the Moon. Both the observed increase in the length of one day of 0.0016 second per century and the observed recession of the Moon of 3 to 4 centimetres per year are understood as consequences of the tides raised on the Earth.

 In Figure 5—>, it has been assumed that the spin axis of mp is perpendicular to the plane of the orbit of ms. If the spin axis is inclined to this plane, the tidal bulge is carried out of the plane as well as ahead of ms. This means that there is a twist, or torque, that changes the direction of the spin axis, so both the magnitude of spin and the direction of the spin axis experience a tidal evolution. The end point of tidal evolution for the spin state of one body of an isolated pair is rotation synchronous with the mean orbital motion with the spin axis perpendicular to the orbit plane. This simple picture is complicated somewhat if other perturbations cause the orbital plane to precess. This precession for the lunar orbit causes its spin axis to be inclined 6°41′ to the orbit plane as the end point of tidal evolution.

      In addition to those of the Earth-Moon pair, numerous other consequences of tidal dissipation and the resulting evolution can be observed in the solar system and elsewhere in the Milky Way Galaxy. For example, all the major and close planetary satellites but one are observed to be rotating synchronously with their orbital motion. The exception is Saturn's (Saturn) satellite Hyperion. Tidal friction has indeed retarded Hyperion's initial spin rate to a value near that of synchronous rotation, but the combination of Hyperion's unusually asymmetric shape and its high orbital eccentricity leads to gravitational (gravitation) torques that make synchronous rotation unstable. As a result, the tides have brought Hyperion to a state where it tumbles chaotically with large changes in the direction and magnitude of its spin on time scales comparable to its orbital period of about 21 days.

      The assembly and maintenance of several orbital resonances among the satellites because of differential tidal expansion of the orbits have also been observed. The orbital resonances among Jupiter's satellites Io, Europa, and Ganymede, where the orbital periods are nearly in the ratio 1:2:4, maintain Io's orbital eccentricity at the value of 0.0041. This rather modest eccentricity causes sufficient variation in the magnitude and direction of Io's enormous tidal bulge to have melted a significant fraction of the satellite through dissipation of tidal energy in spite of Io's synchronous rotation. As a result, Io is the most volcanically active body in the solar system. The orbital eccentricity would normally be damped to zero by this large dissipation, but the orbital resonances with Europa and Ganymede prevent this from happening.

      The distant dwarf planet Pluto and its satellite Charon have almost certainly reached the ultimate end point where further tidal evolution has ceased altogether (the tiny tides raised by the Sun and other planets being neglected). In this state the orbit is circular, with both bodies rotating synchronously with the orbital motion and both spin axes perpendicular to the orbital plane.

      The spin of the planet Mercury has been slowed by tides raised by the Sun to a final state where the spin angular velocity is exactly 1.5 times the orbital mean motion. This state is stable against further change because Mercury's high orbital eccentricity (0.206) allows restoring torques on the permanent (nontidal) axial asymmetry of the planet, which keeps the longest equatorial axis aligned with the direction to the Sun at perihelion. The tidal reduction of Mercury's average eccentricity (near 0.2) will cause insufficient change during the remaining lifetime of the Sun to disrupt this spin-orbit resonance. Finally, many close binary stars are observed to have circular orbits and synchronized spins—an example of tidal evolution elsewhere in the Milky Way Galaxy.

Stanton J. Peale

Additional Reading
Modern introductory treatments and discussions of some advanced techniques and classic developments include J.M.A. Danby, Fundamentals of Celestial Mechanics, 2nd ed., rev. and enlarged (1988); Dirk Brouwer and Gerald M. Clemence, Methods of Celestial Mechanics (1961); and Henry Crozier Keating Plummer, An Introductory Treatise on Dynamical Astronomy (1918, reprinted 1960). Orbital resonances are discussed in two review articles by S.J. Peale: “Orbital Resonances in Solar-System,” Annual Review of Astronomy and Astrophysics, 14:215–246 (1976), and “Orbital Resonances, Unusual Configurations, and Exotic Rotation States Among Planetary Satellites,” in Joseph A. Burns and Mildred Shapley Matthews (eds.), Satellites (1986), pp. 159–223. Current practice in solving the n-body problem on computers is given in the introduction to a paper by Lars Hernquist, “Performance Characteristics of Tree Codes,” The Astrophysical Journal: Supplement Series, 64(4):715–734 (August 1987). An introduction to modern dynamics involving chaos and an introduction to algebraic maps is given by Michel Henon, “Numerical Exploration of Hamiltonian Systems,” in Gérard Iooss, Robert H.G. Helleman, and Raymond Stora (eds.), Chaotic Behaviour in Deterministic Systems (1983), pp. 54–170. Readable accounts of examples of chaotic dynamics in celestial mechanics are found in two articles by Jack Wisdom: “Chaotic Dynamics in the Solar-System,” Icarus, 72(2):241–275 (1987), and “Chaotic Behaviour in the Solar System,” in M.V. Berry, I.C. Percival, and N.O. Weiss (eds.), Dynamical Chaos (1987), pp. 109–129. A simple discussion of tides and tidal evolution is given by S.J. Peale, “Consequences of Tidal Evolution,” in Margaret G. Kivelson (ed.), The Solar System: Observations and Interpretations (1986), pp. 275–288. Advanced discussions of tidal evolution analysis as applied to the Earth are given by Kurt Lambeck, The Earth's Variable Rotation: Geophysical Causes and Consequences (1980). Stanton J. Peale

* * *


Universalium. 2010.

Игры ⚽ Поможем сделать НИР

Look at other dictionaries:

  • Celestial mechanics — is the branch of astrophysics that deals with the motions of celestial objects. The field applies principles of physics, historically classical mechanics, to astronomical objects such as stars and planets to produce ephemeris data. Orbital… …   Wikipedia

  • celestial mechanics — pl.n. the branch of astronomy that deals with the motion and positions of celestial objects in orbit …   English World dictionary

  • celestial mechanics — dangaus mechanika statusas T sritis fizika atitikmenys: angl. celestial mechanics vok. Himmelsmechanik, f rus. небесная механика, f pranc. mécanique céleste, f …   Fizikos terminų žodynas

  • celestial mechanics — noun The dynamics of celestial bodies subject to mutual gravitational attraction …   Wiktionary

  • celestial mechanics — plural noun [treated as sing.] the branch of astronomy concerned with the calculation of the motions of celestial objects such as planets …   English new terms dictionary

  • celestial mechanics — celes′tial mechan′ics n. astron. the branch of astronomy that applies the laws of dynamics and gravitation to the motions of heavenly bodies …   From formal English to slang

  • Celestial Mechanics —    A branch of astronomy which studies the motions and gravitational influences of the members of a solar system …   The writer's dictionary of science fiction, fantasy, horror and mythology

  • celestial mechanics — noun the branch of astronomy concerned with the application of Newton s laws of motion to the motions of heavenly bodies • Hypernyms: ↑astronomy, ↑uranology …   Useful english dictionary

  • mechanics — /meuh kan iks/, n. 1. (used with a sing. v.) the branch of physics that deals with the action of forces on bodies and with motion, comprised of kinetics, statics, and kinematics. 2. (used with a sing. v.) the theoretical and practical application …   Universalium

  • Celestial spheres — Not to be confused with celestial sphere. For other uses, see Celestial (disambiguation). Geocentric celestial spheres; Peter Apian s Cosmographia (Antwerp, 1539) …   Wikipedia

Share the article and excerpts

Direct link
Do a right-click on the link above
and select “Copy Link”