Surface gravity

Last updated

The surface gravity, g, of an astronomical object is the gravitational acceleration experienced at its surface at the equator, including the effects of rotation. The surface gravity may be thought of as the acceleration due to gravity experienced by a hypothetical test particle which is very close to the object's surface and which, in order not to disturb the system, has negligible mass. For objects where the surface is deep in the atmosphere and the radius not known, the surface gravity is given at the 1 bar pressure level in the atmosphere.

Contents

Surface gravity is measured in units of acceleration, which, in the SI system, are meters per second squared. It may also be expressed as a multiple of the Earth's standard surface gravity, which is equal to [1]

g = 9.80665 m/s2

In astrophysics, the surface gravity may be expressed as log g, which is obtained by first expressing the gravity in cgs units, where the unit of acceleration and surface gravity is centimeters per second squared (cm/s2), and then taking the base-10 logarithm of the cgs value of the surface gravity. [2] Therefore, the surface gravity of Earth could be expressed in cgs units as 980.665 cm/s2, and then taking the base-10 logarithm ("log g") of 980.665, and we get 2.992 as "log g".

The surface gravity of a white dwarf is very high, and of a neutron star even higher. A white dwarf's surface gravity is around 100,000 g (106 m/s2) whilst the neutron star's compactness gives it a surface gravity of up to 7×1012 m/s2 with typical values of order 1012 m/s2 (that is more than 1011 times that of Earth). One measure of such immense gravity is that neutron stars have an escape velocity of around 100,000 km/s, about a third of the speed of light. For black holes, the surface gravity must be calculated relativistically.

Relationship of surface gravity to mass and radius

Surface gravity of various
Solar System bodies [3]
(1 g = 9.80665 m/s2, the average surface gravitational acceleration on Earth)
NameSurface gravity
Sun 28.02 g
Mercury 0.377 g
Venus 0.905 g
Earth 1 g (midlatitudes)
Moon 0.165 7 g (average)
Mars 0.379 g (midlatitudes)
Phobos 0.000 581 g
Deimos 0.000 306 g
Pallas 0.022 g (equator)
Vesta 0.025 g (equator)
Ceres 0.029 g
Jupiter 2.528 g (midlatitudes)
Io 0.183 g
Europa 0.134 g
Ganymede 0.146 g
Callisto 0.126 g
Saturn 1.065 g (midlatitudes)
Mimas 0.006 48 g
Enceladus 0.011 5 g
Tethys 0.014 9 g
Dione 0.023 7 g
Rhea 0.026 9 g
Titan 0.138 g
Iapetus 0.022 8 g
Phoebe 0.003 9–0.005 1 g
Uranus 0.886 g (equator)
Miranda 0.007 9 g
Ariel 0.025 4 g
Umbriel 0.023 g
Titania 0.037 2 g
Oberon 0.036 1 g
Neptune 1.137 g (midlatitudes)
Proteus 0.007 g
Triton 0.079 4 g
Pluto 0.063 g
Charon 0.029 4 g
Eris 0.084 g
Haumea 0.0247 g (equator)
67P-CG 0.000 017 g

In the Newtonian theory of gravity, the gravitational force exerted by an object is proportional to its mass: an object with twice the mass produces twice as much force. Newtonian gravity also follows an inverse square law, so that moving an object twice as far away divides its gravitational force by four, and moving it ten times as far away divides it by 100. This is similar to the intensity of light, which also follows an inverse square law: with relation to distance, light becomes less visible. Generally speaking, this can be understood as geometric dilution corresponding to point-source radiation into three-dimensional space.

A large object, such as a planet or star, will usually be approximately round, approaching hydrostatic equilibrium (where all points on the surface have the same amount of gravitational potential energy). On a small scale, higher parts of the terrain are eroded, with eroded material deposited in lower parts of the terrain. On a large scale, the planet or star itself deforms until equilibrium is reached. [4] For most celestial objects, the result is that the planet or star in question can be treated as a near-perfect sphere when the rotation rate is low. However, for young, massive stars, the equatorial azimuthal velocity can be quite high—up to 200 km/s or more—causing a significant amount of equatorial bulge. Examples of such rapidly rotating stars include Achernar, Altair, Regulus A and Vega.

The fact that many large celestial objects are approximately spheres makes it easier to calculate their surface gravity. According to the shell theorem, the gravitational force outside a spherically symmetric body is the same as if its entire mass were concentrated in the center, as was established by Sir Isaac Newton. [5] Therefore, the surface gravity of a planet or star with a given mass will be approximately inversely proportional to the square of its radius, and the surface gravity of a planet or star with a given average density will be approximately proportional to its radius. For example, the recently discovered planet, Gliese 581 c, has at least 5 times the mass of Earth, but is unlikely to have 5 times its surface gravity. If its mass is no more than 5 times that of the Earth, as is expected, [6] and if it is a rocky planet with a large iron core, it should have a radius approximately 50% larger than that of Earth. [7] [8] Gravity on such a planet's surface would be approximately 2.2 times as strong as on Earth. If it is an icy or watery planet, its radius might be as large as twice the Earth's, in which case its surface gravity might be no more than 1.25 times as strong as the Earth's. [8]

These proportionalities may be expressed by the formula:

where g is the surface gravity of an object, expressed as a multiple of the Earth's, m is its mass, expressed as a multiple of the Earth's mass (5.976×1024 kg) and r its radius, expressed as a multiple of the Earth's (mean) radius (6,371 km). [9] For instance, Mars has a mass of 6.4185×1023 kg = 0.107 Earth masses and a mean radius of 3,390 km = 0.532 Earth radii. [10] The surface gravity of Mars is therefore approximately

times that of Earth. Without using the Earth as a reference body, the surface gravity may also be calculated directly from Newton's law of universal gravitation, which gives the formula

where M is the mass of the object, r is its radius, and G is the gravitational constant. If we let ρ = M/V denote the mean density of the object, we can also write this as

so that, for fixed mean density, the surface gravity g is proportional to the radius r.

Since gravity is inversely proportional to the square of the distance, a space station 400 km above the Earth feels almost the same gravitational force as we do on the Earth's surface. A space station does not plummet to the ground because it is in an orbit around the Earth.

Gas giants

For gas giant planets such as Jupiter, Saturn, Uranus, and Neptune, the surface gravity is given at the 1 bar pressure level in the atmosphere. [11]

Non-spherically symmetric objects

Most real astronomical objects are not perfectly spherically symmetric. One reason for this is that they are often rotating, which means that they are affected by the combined effects of gravitational force and centrifugal force. This causes stars and planets to be oblate, which means that their surface gravity is smaller at the equator than at the poles. This effect was exploited by Hal Clement in his SF novel Mission of Gravity , dealing with a massive, fast-spinning planet where gravity was much higher at the poles than at the equator.

To the extent that an object's internal distribution of mass differs from a symmetric model, we may use the measured surface gravity to deduce things about the object's internal structure. This fact has been put to practical use since 1915–1916, when Roland Eötvös's torsion balance was used to prospect for oil near the city of Egbell (now Gbely, Slovakia.) [12] :1663 [13] :223 In 1924, the torsion balance was used to locate the Nash Dome oil fields in Texas. [13] :223

It is sometimes useful to calculate the surface gravity of simple hypothetical objects which are not found in nature. The surface gravity of infinite planes, tubes, lines, hollow shells, cones, and even more unrealistic structures may be used to provide insights into the behavior of real structures.

Black holes

In relativity, the Newtonian concept of acceleration turns out not to be clear cut. For a black hole, which must be treated relativistically, one cannot define a surface gravity as the acceleration experienced by a test body at the object's surface because there is no surface. This is because the acceleration of a test body at the event horizon of a black hole turns out to be infinite in relativity. Because of this, a renormalized value is used that corresponds to the Newtonian value in the non-relativistic limit. The value used is generally the local proper acceleration (which diverges at the event horizon) multiplied by the gravitational time dilation factor (which goes to zero at the event horizon). For the Schwarzschild case, this value is mathematically well behaved for all non-zero values of r and M.

When one talks about the surface gravity of a black hole, one is defining a notion that behaves analogously to the Newtonian surface gravity, but is not the same thing. In fact, the surface gravity of a general black hole is not well defined. However, one can define the surface gravity for a black hole whose event horizon is a Killing horizon.

The surface gravity of a static Killing horizon is the acceleration, as exerted at infinity, needed to keep an object at the horizon. Mathematically, if is a suitably normalized Killing vector, then the surface gravity is defined by

where the equation is evaluated at the horizon. For a static and asymptotically flat spacetime, the normalization should be chosen so that as , and so that . For the Schwarzschild solution, we take to be the time translation Killing vector , and more generally for the Kerr–Newman solution we take , the linear combination of the time translation and axisymmetry Killing vectors which is null at the horizon, where is the angular velocity.

Schwarzschild solution

Since is a Killing vector implies . In coordinates . Performing a coordinate change to the advanced Eddington–Finklestein coordinates causes the metric to take the form

Under a general change of coordinates the Killing vector transforms as giving the vectors and

Considering the b = entry for gives the differential equation

Therefore, the surface gravity for the Schwarzschild solution with mass is ( in SI units). [14]

Kerr solution

The surface gravity for the uncharged, rotating black hole is, simply

where is the Schwarzschild surface gravity, and is the spring constant of the rotating black hole. is the angular velocity at the event horizon. This expression gives a simple Hawking temperature of . [15]

Kerr–Newman solution

The surface gravity for the Kerr–Newman solution is

where is the electric charge, is the angular momentum, we define to be the locations of the two horizons and .

Dynamical black holes

Surface gravity for stationary black holes is well defined. This is because all stationary black holes have a horizon that is Killing. [16] Recently there has been a shift towards defining the surface gravity of dynamical black holes whose spacetime does not admit a timelike Killing vector (field). [17] Several definitions have been proposed over the years by various authors, such as peeling surface gravity and Kodama surface gravity. [18] As of current, there is no consensus or agreement on which definition, if any, is correct. [19] Semiclassical results indicate that the peeling surface gravity is ill-defined for transient objects formed in finite time of a distant observer. [20]

Related Research Articles

<span class="mw-page-title-main">Mass</span> Amount of matter present in an object

Mass is an intrinsic property of a body. It was traditionally believed to be related to the quantity of matter in a body, until the discovery of the atom and particle physics. It was found that different atoms and different elementary particles, theoretically with the same amount of matter, have nonetheless different masses. Mass in modern physics has multiple definitions which are conceptually distinct, but physically equivalent. Mass can be experimentally defined as a measure of the body's inertia, meaning the resistance to acceleration when a net force is applied. The object's mass also determines the strength of its gravitational attraction to other bodies.

<span class="mw-page-title-main">Tidal force</span> A gravitational effect also known as the differential force and the perturbing force

The tidal force or tide-generating force is a gravitational effect that stretches a body along the line towards and away from the center of mass of another body due to spatial variations in strength in gravitational field from the other body. It is responsible for the tides and related phenomena, including solid-earth tides, tidal locking, breaking apart of celestial bodies and formation of ring systems within the Roche limit, and in extreme cases, spaghettification of objects. It arises because the gravitational field exerted on one body by another is not constant across its parts: the nearer side is attracted more strongly than the farther side. The difference is positive in the near side and negative in the far side, which causes a body to get stretched. Thus, the tidal force is also known as the differential force, residual force, or secondary effect of the gravitational field.

<span class="mw-page-title-main">Escape velocity</span> Concept in celestial mechanics

In celestial mechanics, escape velocity or escape speed is the minimum speed needed for an object to escape from contact with or orbit of a primary body, assuming:

<span class="mw-page-title-main">Gravitational binding energy</span> Minimum energy to remove a system from a gravitationally bound state

The gravitational binding energy of a system is the minimum energy which must be added to it in order for the system to cease being in a gravitationally bound state. A gravitationally bound system has a lower gravitational potential energy than the sum of the energies of its parts when these are completely separated—this is what keeps the system aggregated in accordance with the minimum total potential energy principle.

<span class="mw-page-title-main">Gravitational field</span> Model in physics

In physics, a gravitational field or gravitational acceleration field is a vector field used to explain the influences that a body extends into the space around itself. A gravitational field is used to explain gravitational phenomena, such as the gravitational force field exerted on another massive body. It has dimension of acceleration (L/T2) and it is measured in units of newtons per kilogram (N/kg) or, equivalently, in meters per second squared (m/s2).

The Schwarzschild radius or the gravitational radius is a physical parameter in the Schwarzschild solution to Einstein's field equations that corresponds to the radius defining the event horizon of a Schwarzschild black hole. It is a characteristic radius associated with any quantity of mass. The Schwarzschild radius was named after the German astronomer Karl Schwarzschild, who calculated this exact solution for the theory of general relativity in 1916.

Newton's law of universal gravitation says that every particle attracts every other particle in the universe with a force that is proportional to the product of their masses and inversely proportional to the square of the distance between their centers. Separated objects attract and are attracted as if all their mass were concentrated at their centers. The publication of the law has become known as the "first great unification", as it marked the unification of the previously described phenomena of gravity on Earth with known astronomical behaviors.

In Einstein's theory of general relativity, the Schwarzschild metric is an exact solution to the Einstein field equations that describes the gravitational field outside a spherical mass, on the assumption that the electric charge of the mass, angular momentum of the mass, and universal cosmological constant are all zero. The solution is a useful approximation for describing slowly rotating astronomical objects such as many stars and planets, including Earth and the Sun. It was found by Karl Schwarzschild in 1916.

<span class="mw-page-title-main">Cavendish experiment</span> Experiment measuring the force of gravity (1797–1798)

The Cavendish experiment, performed in 1797–1798 by English scientist Henry Cavendish, was the first experiment to measure the force of gravity between masses in the laboratory and the first to yield accurate values for the gravitational constant. Because of the unit conventions then in use, the gravitational constant does not appear explicitly in Cavendish's work. Instead, the result was originally expressed as the specific gravity of Earth, or equivalently the mass of Earth. His experiment gave the first accurate values for these geophysical constants.

Photoevaporation is the process where energetic radiation ionises gas and causes it to disperse away from the ionising source. The term is typically used in an astrophysical context where ultraviolet radiation from hot stars acts on clouds of material such as molecular clouds, protoplanetary disks, or planetary atmospheres.

In physics and astronomy, the Reissner–Nordström metric is a static solution to the Einstein–Maxwell field equations, which corresponds to the gravitational field of a charged, non-rotating, spherically symmetric body of mass M. The analogous solution for a charged, rotating body is given by the Kerr–Newman metric.

In physics, a Killing horizon is a geometrical construct used in general relativity and its generalizations to delineate spacetime boundaries without reference to the dynamic Einstein field equations. Mathematically a Killing horizon is a null hypersurface defined by the vanishing of the norm of a Killing vector field. It can also be defined as a null hypersurface generated by a Killing vector, which in turn is null at that surface.

<span class="mw-page-title-main">Gravity train</span> Theoretical means of transportation

A gravity train is a theoretical means of transportation for purposes of commuting between two points on the surface of a sphere, by following a straight tunnel connecting the two points through the interior of the sphere.

<span class="mw-page-title-main">Gravity of Earth</span>

The gravity of Earth, denoted by g, is the net acceleration that is imparted to objects due to the combined effect of gravitation and the centrifugal force . It is a vector quantity, whose direction coincides with a plumb bob and strength or magnitude is given by the norm .

<span class="mw-page-title-main">Proper acceleration</span> Physical acceleration experienced by an object

In relativity theory, proper acceleration is the physical acceleration experienced by an object. It is thus acceleration relative to a free-fall, or inertial, observer who is momentarily at rest relative to the object being measured. Gravitation therefore does not cause proper acceleration, because the same gravity acts equally on the inertial observer. As a consequence, all inertial observers always have a proper acceleration of zero.

The two-body problem in general relativity is the determination of the motion and gravitational field of two bodies as described by the field equations of general relativity. Solving the Kepler problem is essential to calculate the bending of light by gravity and the motion of a planet orbiting its sun. Solutions are also used to describe the motion of binary stars around each other, and estimate their gradual loss of energy through gravitational radiation.

<span class="mw-page-title-main">Gravitoelectromagnetism</span> Analogies between Maxwells and Einsteins field equations

Gravitoelectromagnetism, abbreviated GEM, refers to a set of formal analogies between the equations for electromagnetism and relativistic gravitation; specifically: between Maxwell's field equations and an approximation, valid under certain conditions, to the Einstein field equations for general relativity. Gravitomagnetism is a widely used term referring specifically to the kinetic effects of gravity, in analogy to the magnetic effects of moving electric charge. The most common version of GEM is valid only far from isolated sources, and for slowly moving test particles.

<span class="mw-page-title-main">Entropic gravity</span> Theory in modern physics that describes gravity as an entropic force

Entropic gravity, also known as emergent gravity, is a theory in modern physics that describes gravity as an entropic force—a force with macro-scale homogeneity but which is subject to quantum-level disorder—and not a fundamental interaction. The theory, based on string theory, black hole physics, and quantum information theory, describes gravity as an emergent phenomenon that springs from the quantum entanglement of small bits of spacetime information. As such, entropic gravity is said to abide by the second law of thermodynamics under which the entropy of a physical system tends to increase over time.

In general relativity, the Vaidya metric describes the non-empty external spacetime of a spherically symmetric and nonrotating star which is either emitting or absorbing null dusts. It is named after the Indian physicist Prahalad Chunnilal Vaidya and constitutes the simplest non-static generalization of the non-radiative Schwarzschild solution to Einstein's field equation, and therefore is also called the "radiating(shining) Schwarzschild metric".

The Ellis drainhole is the earliest-known complete mathematical model of a traversable wormhole. It is a static, spherically symmetric solution of the Einstein vacuum field equations augmented by inclusion of a scalar field minimally coupled to the geometry of space-time with coupling polarity opposite to the orthodox polarity :

References

  1. Taylor, Barry N., ed. (2001). The International System of Units (SI) (PDF). United States Department of Commerce: National Institute of Standards and Technology. p. 29. Retrieved 2012-03-08.{{cite book}}: |work= ignored (help)
  2. Smalley, B. (13 July 2006). "The Determination of Teff and log g for B to G stars". Keele University . Retrieved 31 May 2007.
  3. Isaac Asimov (1978). The Collapsing Universe. Corgi. p. 44. ISBN   978-0-552-10884-3.
  4. "Why is the Earth round?". Ask A Scientist. Argonne National Laboratory, Division of Educational Programs. Archived from the original on 21 September 2008.
  5. Book I, §XII, pp. 218–226, Newton's Principia: The Mathematical Principles of Natural Philosophy, Sir Isaac Newton, tr. Andrew Motte, ed. N. W. Chittenden. New York: Daniel Adee, 1848. First American edition.
  6. Astronomers Find First Earth-like Planet in Habitable Zone Archived 2009-06-17 at the Wayback Machine , ESO 22/07, press release from the European Southern Observatory, April 25, 2007
  7. Udry, Stéphane; Bonfils, Xavier; Delfosse, Xavier; Forveille, Thierry; Mayor, Michel; Perrier, Christian; Bouchy, François; Lovis, Christophe; Pepe, Francesco; Queloz, Didier; Bertaux, Jean-Loup (2007). "The HARPS search for southern extra-solar planets XI. Super-Earths (5 and 8 ME) in a 3-planet system" (PDF). Astronomy & Astrophysics. 469 (3): L43–L47. arXiv: 0704.3841 . Bibcode:2007A&A...469L..43U. doi:10.1051/0004-6361:20077612. S2CID   119144195. Archived from the original (PDF) on October 8, 2010.
  8. 1 2 Valencia, Diana; Sasselov, Dimitar D; O'Connell, Richard J (2007). "Detailed Models of super-Earths: How well can we infer bulk properties?". The Astrophysical Journal. 665 (2): 1413–1420. arXiv: 0704.3454 . Bibcode:2007ApJ...665.1413V. doi:10.1086/519554. S2CID   15605519.
  9. 2.7.4 Physical properties of the Earth, web page, accessed on line May 27, 2007.
  10. Mars Fact Sheet, web page at NASA NSSDC, accessed May 27, 2007.
  11. "Planetary Fact Sheet Notes".
  12. Li, Xiong; Götze, Hans-Jürgen (2001). "Ellipsoid, geoid, gravity, geodesy, and geophysics". Geophysics. 66 (6): 1660–1668. Bibcode:2001Geop...66.1660L. doi:10.1190/1.1487109.
  13. 1 2 Prediction by Eötvös' torsion balance data in Hungary Archived 2007-11-28 at the Wayback Machine , Gyula Tóth, Periodica Polytechnica Ser. Civ. Eng.46, #2 (2002), pp. 221–229.
  14. Raine, Derek J.; Thomas, Edwin George (2010). Black Holes: An Introduction (illustrated ed.). Imperial College Press. p. 44. ISBN   978-1-84816-382-9. Extract of page 44
  15. Good, Michael; Yen Chin Ong (February 2015). "Are Black Holes Springlike?". Physical Review D. 91 (4): 044031. arXiv: 1412.5432 . Bibcode:2015PhRvD..91d4031G. doi:10.1103/PhysRevD.91.044031. S2CID   117749566.
  16. Wald, Robert (1984). General Relativity . University Of Chicago Press. ISBN   978-0-226-87033-5.
  17. A. B. Nielsen; J. H. Yoon (2008). "Dynamical Surface Gravity". Classical and Quantum Gravity. 25 (8): 085010. arXiv: 0711.1445 . Bibcode:2008CQGra..25h5010N. doi:10.1088/0264-9381/25/8/085010. S2CID   15438397.
  18. H. Kodama (1980). "Conserved Energy Flux for the Spherically Symmetric System and the Backreaction Problem in the Black Hole Evaporation". Progress of Theoretical Physics. 63 (4): 1217. Bibcode:1980PThPh..63.1217K. doi:10.1143/PTP.63.1217. S2CID   122827579.
  19. Pielahn, Mathias; G. Kunstatter; A. B. Nielsen (November 2011). "Dynamical surface gravity in spherically symmetric black hole formation". Physical Review D. 84 (10): 104008(11). arXiv: 1103.0750 . Bibcode:2011PhRvD..84j4008P. doi:10.1103/PhysRevD.84.104008. S2CID   119015033.
  20. R. B. Mann; S. Murk; D. R. Terno (2022). "Surface gravity and the information loss problem". Physical Review D. 105 (12): 124032. arXiv: 2109.13939 . Bibcode:2022PhRvD.105l4032M. doi:10.1103/PhysRevD.105.124032. S2CID   249799593.