Fact-checked by Grok 2 weeks ago

Gravitational acceleration

Gravitational acceleration is the acceleration experienced by a free-falling object due to the gravitational attraction of a massive body, independent of the object's mass in a vacuum. On Earth, it is often denoted as g and arises primarily from the planet's gravitational field. Near Earth's surface, this value varies slightly by location but is standardized at exactly 9.80665 m/s² for reference purposes in physics and engineering. In general, gravitational acceleration follows from Isaac Newton's law of universal gravitation, which states that the gravitational force between two masses is proportional to the product of their masses and inversely proportional to the square of the distance between their centers. For an object near a spherical body's surface, it can be expressed as g = GM / r², where G is the gravitational constant (6.67430 × 10⁻¹¹ m³ kg⁻¹ s⁻²), M is the body's mass, and r is the distance from its center. For Earth, M ≈ 5.973 × 10²⁴ kg and equatorial r ≈ 6.378 × 10⁶ m, yielding an average surface value of approximately 9.8 m/s²—meaning a dropped object increases its speed by about 9.8 meters per second every second. Notably, g is not uniform across Earth: it is stronger at the poles (around 9.832 m/s²) due to closer proximity to Earth's center and the absence of centrifugal effects from rotation, and weaker at the equator (about 9.780 m/s²) because of the planet's oblate shape and rotational bulge. Altitude also influences g, decreasing with height above sea level as the inverse-square law predicts, though this effect is small near the surface (e.g., a 1 km increase reduces g by about 0.03%). These variations are critical in fields like geodesy, satellite orbits, and precise measurements of mass using gravimeters.

Classical Description

Definition and Basic Concepts

Gravitational acceleration, often denoted as g, is the acceleration imparted to a freely falling object in a gravitational field, arising solely from the attractive force of gravity and assuming negligible air resistance. This acceleration represents a key concept in kinematics, where acceleration is defined as the rate of change of velocity with respect to time, typically measured in meters per second squared (m/s²) in the International System of Units (SI). In the context of Earth's gravity, g governs the motion of objects in free fall, causing them to speed up uniformly toward the planet's center. Near Earth's surface, gravitational acceleration is represented as a vector quantity, \vec{g}, pointing downward toward the Earth's center, with a magnitude of approximately 9.80665 m/s² at sea level under standard conditions. The vector nature of \vec{g} emphasizes both its direction and magnitude, which can vary slightly depending on location due to local gravitational influences, though the standard value serves as a global reference. An alternative historical unit for gravitational acceleration is the gal (named after Galileo Galilei), where 1 gal equals 0.01 m/s², making Earth's standard g equivalent to about 980.665 gal. It is important to distinguish gravitational acceleration from weight: while g is an acceleration independent of the object's mass, weight (W) is the gravitational force on an object, given by W = m g, where m is the mass. Thus, all objects in free fall experience the same g, regardless of mass, leading to equal accelerations for feathers and hammers in a vacuum, as famously demonstrated in experiments. This local manifestation of gravity aligns with Newton's law of universal gravitation, which explains the underlying attractive force between masses.

Relation to Newton's Law of Universal Gravitation

Newton's law of universal gravitation states that every particle in the universe attracts every other particle with a force that is directly proportional to the product of their masses and inversely proportional to the square of the distance between their centers. This force F is given by the equation F = G \frac{m_1 m_2}{r^2}, where m_1 and m_2 are the masses of the two particles, r is the distance between their centers, and G is the gravitational constant, with a value of $6.67430 \times 10^{-11} \, \mathrm{m}^3 \mathrm{kg}^{-1} \mathrm{s}^{-2}. For gravitational acceleration near a massive body, such as Earth, consider an object of mass m at a distance r from the center of a spherical body of mass M. The gravitational force on the object is F = G \frac{M m}{r^2}. By Newton's second law, this force equals m g, where g is the acceleration due to gravity. Thus, g = \frac{G M}{r^2}, showing that gravitational acceleration emerges as a special case of the universal law for objects near the surface of a much more massive body, where the test mass m cancels out. This derivation assumes the central body is spherically symmetric and the distance r is much larger than the object's size, treating it as a point mass. Treating Earth as a uniform sphere, this formula yields an approximate value of g \approx 9.8 \, \mathrm{m/s}^2 at the surface, where r is Earth's radius and M is its mass. The inverse square dependence implies that g decreases with increasing distance from the center of the body, halving when r doubles, for instance. Isaac Newton formulated this law in his Philosophiæ Naturalis Principia Mathematica, published in 1687, thereby unifying the force causing objects to fall on Earth with the gravitational attraction governing celestial motions, such as planetary orbits.

Variations and Influences

Effects on Earth

Gravitational acceleration on Earth, denoted as g, exhibits variations across the planet's surface primarily due to its oblate spheroid shape, rotational motion, and local geological features. These effects result in an effective g that differs from the idealized value derived from Newton's law of universal gravitation, which assumes a spherical, non-rotating body. The most prominent variation occurs with latitude, where g is stronger at the poles, approximately 9.832 m/s², compared to the equator at about 9.780 m/s². This difference, roughly 0.5%, arises from two main factors: the centrifugal force due to Earth's rotation, which reduces the effective g more significantly at the equator, and the equatorial bulge caused by rotation, which increases the distance from the planet's center at lower latitudes, thereby weakening gravitational pull. The centrifugal effect specifically contributes a reduction in effective g given by \omega^2 r \cos^2 \phi, where \omega is Earth's angular velocity (approximately $7.292 \times 10^{-5} rad/s), r is the Earth's radius at the location, and \phi is the latitude. This outward acceleration is maximal at the equator (\phi = 0^\circ, \cos \phi = 1) and zero at the poles (\phi = 90^\circ), accounting for about 0.3% of the total variation in g. Altitude also influences g, with values decreasing as height above sea level increases because of the greater distance from Earth's center of mass. The free-air correction approximates this change as a reduction of about $3.086 \times 10^{-6} m/s² per meter of elevation, or equivalently 0.3086 mGal/m, reflecting the inverse-square law of gravitation for small heights. To model the latitude-dependent variation at sea level, the International Gravity Formula (IGF), adopted in 1930 by the International Association of Geodesy, provides an empirical expression: g(\phi) = 9.780327 \left(1 + 0.0053024 \sin^2 \phi - 0.0000058 \sin^2 2\phi \right) \ \text{m/s}^2 This formula incorporates both gravitational and centrifugal effects for the reference ellipsoid, yielding values consistent with observed data across latitudes. Superimposed on these global patterns are local geological anomalies caused by subsurface density variations, such as denser rock formations or mineral deposits, which can increase or decrease g by small amounts. For instance, positive anomalies of up to +0.2 mGal may occur over mountainous regions due to the additional mass of elevated terrain in free-air measurements, while sedimentary basins often produce negative anomalies from lower-density materials. These anomalies, typically on the order of milligals, are crucial for geophysical prospecting and understanding crustal structure.

Factors Affecting Measurement

Accurate measurement of gravitational acceleration is influenced by several environmental factors that introduce temporal and spatial variations. Tidal effects, primarily from the gravitational pull of the Moon and Sun, cause periodic changes in Earth's gravity field, with maximum variations reaching up to 0.3 mGal over a tidal cycle, including contributions from solid Earth deformation of about 0.04 mGal. These effects exhibit semi-diurnal cycles, typically twice per lunar day, complicating precise determinations unless modeled and corrected. Atmospheric and hydrological influences further perturb measurements through loading effects. Variations in air pressure alter gravity by approximately -0.3 μGal per mbar due to changes in atmospheric mass distribution, requiring corrections based on local barometric data. Similarly, hydrological loading from water bodies, soil moisture, and ice masses induces gravity changes; for instance, seasonal snow or groundwater fluctuations can produce signals on the order of several μGal, necessitating integration of global or regional hydrological models for accurate subtraction. Instrumental errors in gravimeters pose significant challenges to precision. In spring-based relative gravimeters, tilt misalignment can introduce errors exceeding 2 μGal if leveling is not calibrated to better than 10 arcseconds, while temperature sensitivity affects spring constants, leading to reading shifts that must be minimized through environmental controls. Drift, arising from mechanical relaxation and aging, causes nonlinear changes over time, often corrected via frequent calibrations or modeling of the instrument's response. To mitigate these issues, calibration against absolute gravimeters is essential, as these instruments achieve precisions of about 2 μGal by directly measuring free-fall acceleration in a vacuum. They serve as references for tying relative measurements to the international gravity datum, ensuring traceability and reducing systematic biases from environmental and instrumental sources. In microgravity environments, such as parabolic aircraft flights or orbital free fall, effective gravitational acceleration approaches 0g for durations of 20-30 seconds per parabola, fundamentally altering measurement setups by eliminating standard weight-based references and requiring inertial or acceleration-based techniques. These conditions, while useful for simulating space, demand specialized adaptations to avoid confounding residual accelerations from vehicle dynamics.

Comparative Values

Across Celestial Bodies

Gravitational acceleration, or surface gravity, on a celestial body is the acceleration experienced by an object at its surface due to the body's gravitational field. For idealized spherical bodies, this is given by the formula g = \frac{GM}{r^2}, where G is the gravitational constant, M is the mass of the body, and r is its radius. This expression arises from Newton's law of universal gravitation applied to the surface, highlighting how surface gravity scales directly with mass and inversely with the square of the radius. The value of surface gravity thus depends strongly on a body's mass and size, with denser objects exhibiting higher acceleration despite smaller radii. For instance, Jupiter, with a mass over 300 times that of Earth but a radius only about 11 times larger, has a surface gravity of 24.79 m/s²—more than 2.5 times Earth's—due to its substantial mass. This underscores that compactness plays a key role in enhancing gravitational pull. For non-spherical bodies, such as oblate spheroids formed by rotation, surface gravity varies with latitude. The equatorial bulge increases the distance from the center, reducing g at the equator compared to the poles, while centrifugal effects further diminish the effective gravity there; this latitude dependence follows from the body's ellipsoidal shape in hydrostatic equilibrium. Surface gravity also relates loosely to escape velocity, v_\mathrm{esc} = \sqrt{2gr}, providing a practical link for estimating the speed needed to break free from a body's gravitational influence, which aids in orbital mechanics calculations without full derivation here. In extreme cases, surface gravity reaches extraordinary levels on highly compact objects. Neutron stars, with masses around 1.4 solar masses compressed into radii of about 10-15 km, exhibit surface gravities on the order of $10^{12} m/s², over 100 billion times Earth's. For black holes in the classical Newtonian limit, gravitational acceleration approaches infinity as one nears the event horizon, though this represents an idealized point-mass approximation beyond which general relativity is required.

Solar System Examples

Gravitational acceleration varies significantly across the Solar System, primarily due to differences in body mass and radius, as governed by Newton's law of universal gravitation. On Earth, the standard value at sea level is 9.807 m/s², serving as a reference for comparisons. The Moon experiences approximately one-sixth of Earth's gravity at 1.625 m/s², resulting from its much lower mass despite a relatively similar density to Earth. At the Sun's photosphere, gravitational acceleration reaches 274 m/s², driven by the star's enormous mass overwhelming its large radius. Among the planets, terrestrial worlds generally have lower values than gas giants, though Venus approaches Earth's due to its similar size and mass. The following table summarizes representative surface gravitational accelerations (equatorial values in m/s²) for major Solar System bodies, derived from NASA's planetary fact sheets and other sources.
BodyGravitational Acceleration (m/s²)
Mercury3.70
Venus8.87
Earth9.807
Moon1.625
Mars3.71
Jupiter24.79
Saturn10.44
Uranus8.87
Neptune11.15
Pluto0.62
Europa1.31
Gas giants like Jupiter exhibit the highest planetary values, exceeding 24 m/s², because their immense masses far outweigh the effect of their large radii, despite lower densities compared to rocky planets. In contrast, icy moons such as Europa have modest accelerations around 1.31 m/s², varying based on their small sizes and compositions, which influence potential habitability and surface processes.

Relativistic Perspective

Equivalence Principle

The weak equivalence principle (WEP) states that all objects, regardless of their mass or composition, undergo the same acceleration in a gravitational field, implying the equality of inertial and gravitational mass. This concept traces its origins to Galileo's 17th-century insight that bodies of different masses fall at the same rate in the absence of air resistance, a universality later incorporated into Newton's law of universal gravitation but without a deeper theoretical foundation. Einstein formalized the WEP in 1907 as a cornerstone of general relativity, elevating it from an empirical observation to a fundamental postulate that all test bodies follow the same trajectories in free fall. Einstein's development of the principle stemmed from what he described as the "happiest thought of my life" in 1907: an observer in free fall experiences no gravitational force, as if in a gravity-free environment. This realization, detailed in his paper "On the Relativity Principle and the Conclusions Drawn from It," bridged the gap between special relativity and a theory of gravitation by positing that the effects of gravity are locally indistinguishable from those of acceleration in a non-inertial reference frame. A classic illustration is the elevator thought experiment: in a sealed elevator accelerating upward at 9.81 m/s² in deep space, a dropped object appears to fall toward the floor just as it would under Earth's gravity; conversely, in free fall within a gravitational field, occupants feel weightless, rendering the scenarios experimentally equivalent on small scales where tidal effects are negligible. Experimental verification of the WEP has progressed from qualitative demonstrations to high-precision measurements. Loránd Eötvös's torsion balance experiments in the late 19th and early 20th centuries first quantitatively tested the equality of gravitational and inertial mass using hanging weights of different materials, achieving a precision of about 10^{-8} in initial setups and improving to 3 \times 10^{-9} by 1922 through refined analyses. Modern ground-based torsion balance tests, such as those by the Eöt-Wash collaboration, have confirmed the principle to 10^{-13} by comparing accelerations of laboratory masses. Atomic interferometry offers even greater sensitivity; for instance, simultaneous interferometers with rubidium isotopes have demonstrated differential accelerations below 10^{-15} g, where g is Earth's gravitational acceleration, using wavepacket interference over meter-scale drops. These results affirm the WEP to extraordinary levels, with no observed deviations. While the WEP holds robustly in classical and relativistic regimes, hypothetical violations arise in some quantum gravity theories, such as those involving spacetime foam or modified dispersion relations at the Planck scale (around 10^{-35} m), where gravitational interactions might differ for particles of varying quantum properties. However, no such violations have been detected experimentally, and tests continue to constrain these models to ever-tighter bounds.

Frame-Dragging and Advanced Effects

In general relativity, frame-dragging, also known as the Lense-Thirring effect, arises when a rotating mass twists the surrounding spacetime, causing a dragging of inertial frames and a consequent alteration to the paths of nearby objects. This effect modifies the effective gravitational acceleration by inducing a small transverse component, with a magnitude of up to approximately $10^{-14} m/s² near Earth's surface due to its rotation. The phenomenon was first predicted in 1918 by Josef Lense and Hans Thirring through an approximate solution to Einstein's field equations for a rotating body. The frame-dragging effect was precisely measured by the Gravity Probe B (GP-B) experiment, launched in 2004 and operational until 2011, which used four superconducting gyroscopes in a polar orbit around Earth to detect spacetime distortions. The mission confirmed the frame-dragging precession at -37.2 ± 7.2 milliarcseconds per year, agreeing with general relativity's prediction of -39.2 milliarcseconds per year to within 19% accuracy. Complementing this, the geodetic precession—resulting from the curvature of spacetime along the orbital path— was measured at -6,601.8 ± 18.3 milliarcseconds per year, matching theory to 0.28% precision and thus confirming general relativity's predictions for these combined effects on gyroscopic motion to high accuracy. In strong gravitational fields near rotating black holes, described by the Kerr metric, frame-dragging becomes pronounced, with spacetime rotation approaching the speed of light at the ergosphere. Here, the effective gravitational acceleration diverges as one approaches the event horizon, where the escape velocity equals the speed of light c, rendering outward motion impossible for massive particles. This limit highlights how relativistic effects dominate, constraining accelerations such that no object can exceed c locally, even as proper acceleration grows unbounded for stationary observers near the horizon. The post-Newtonian approximation provides a perturbative framework to incorporate these relativistic corrections to Newtonian gravity for weak fields and slow motions. The relativistic gravitational acceleration is given by \mathbf{g}_\text{rel} = \mathbf{g}_\text{Newton} \left(1 + \text{corrections}\right), where key terms include factors like \frac{2GM}{c^2 r} that account for spacetime curvature effects, as evidenced in the anomalous precession of Mercury's orbit, which advances by 43 arcseconds per century beyond Newtonian predictions. Observational evidence for these relativistic modifications to gravitational acceleration is routinely applied in the Global Positioning System (GPS), where satellites at approximately 20,200 km altitude experience variations in effective g due to general relativity. These necessitate clock adjustments for gravitational time dilation and velocity effects, resulting in a net daily drift of about 38 microseconds if uncorrected, ensuring positional accuracy to within meters.

Determination and Applications

Historical Methods

The measurement of gravitational acceleration, denoted as g, has evolved significantly from qualitative observations to precise quantitative techniques. In the late 16th century, Galileo Galilei conducted pioneering experiments to demonstrate that objects fall with constant acceleration independent of their mass. Around 1590, he reportedly dropped objects of different masses from the Leaning Tower of Pisa, observing that they struck the ground simultaneously, challenging Aristotelian notions of motion. To study this more accurately, Galileo used inclined planes, rolling balls down ramps of varying angles and timing their descent with a water clock; these experiments revealed that the acceleration along the plane was proportional to the sine of the angle, allowing extrapolation to free fall and establishing g as uniform. By the mid-17th century, the pendulum emerged as a key tool for measuring g. In 1656, Christiaan Huygens invented the pendulum clock and derived the relationship between its period and gravitational acceleration. For a simple pendulum of length L, the period T is given by T = 2\pi \sqrt{\frac{L}{g}}, enabling g to be computed as g = 4\pi^2 L / T^2 from measurements of L and T. This method provided relative values of g at different locations, with accuracies improved to about 0.1% by careful construction of cycloidal cheeks to ensure isochronous swings. In the 18th century, efforts focused on linking g to Earth's density and shape. The Schiehallion experiment, conducted by Nevil Maskelyne in 1774 on the Scottish mountain Schiehallion, measured the deflection of a plumb line due to the mountain's gravitational pull using a zenith sector telescope. This deflection, about 11 arcseconds toward the mountain, allowed calculation of Schiehallion's density and, by extension, Earth's mean density, providing indirect insight into variations in g. Complementing this, Henry Cavendish's 1798 torsion balance experiment directly measured the gravitational constant G between lead spheres, yielding G ≈ 6.74 × 10^{-11} m³ kg^{-1} s^{-2}; since g = G M / R^2 for Earth, this enabled derivation of Earth's mass and confirmed g's magnitude. The 19th century brought refinements in pendulum techniques for higher precision. In 1817, Henry Kater developed the reversible pendulum, a compound device with two knife-edge pivots and adjustable bobs that could swing about either pivot. By balancing the periods when suspended from each end to approximately 2 seconds, systematic errors from pivot mass were minimized, achieving g measurements accurate to 0.01% or better; Kater's value at London was g ≈ 9.8118 m/s². This instrument became a standard for absolute gravimetry until the mid-20th century. In the early 20th century, absolute measurements shifted toward free-fall methods to avoid pendulum uncertainties. Vacuum drop towers and tubes were employed to measure the time of fall over known distances, eliminating air resistance; experiments in the 1930s using such setups achieved accuracies around 0.1%, paving the way for modern interferometric techniques.

Modern Techniques and Uses

Modern techniques for measuring gravitational acceleration have advanced significantly, enabling high-precision determinations essential for scientific and practical applications. Absolute gravimeters, such as the FG5 series developed by Micro-g LaCoste, utilize interferometric methods involving a falling corner cube retro-reflector in a vacuum chamber to measure free-fall distance with laser interferometry. These instruments achieve a precision of approximately 2 μGal (where 1 μGal = 10^{-8} m/s²) through multiple drops, providing absolute values of g independent of calibration against reference sites. Relative gravimeters complement these by detecting variations in g across locations; spring-based models like the LaCoste-Romberg use a zero-length spring suspension to sense small changes in gravitational force on a proof mass, suitable for field surveys with resolutions down to 10-20 μGal. Superconducting gravimeters, employing magnetic levitation of a niobium sphere in a cryogenic environment, offer continuous monitoring with noise levels below 0.1 μGal over long periods, ideal for tidal and geophysical studies. Satellite gravimetry represents a global-scale approach, with missions like GRACE (2002-2017), its follow-on GRACE-FO (2018–present), and GOCE (2009-2013) mapping Earth's gravity field by analyzing satellite orbit perturbations caused by mass variations. GRACE, using twin satellites to measure inter-satellite distance changes via microwave ranging, resolved temporal gravity variations to about 10^{-5} m/s² over spatial scales of 300-400 km, while GOCE's electrostatic accelerometers compensated for non-gravitational forces in a low-Earth orbit, achieving static field resolutions of similar magnitude at finer 100-150 km scales. These techniques have revolutionized geophysics; for instance, gravity anomalies detected by ground and satellite surveys guide resource exploration by identifying subsurface density contrasts associated with oil and mineral deposits, such as negative anomalies over low-density sedimentary basins. In earthquake monitoring, microgravity measurements linked to crustal strain—where dilatation causes subtle gravity decreases of 1-10 μGal—aid in assessing preseismic deformation patterns. Beyond Earth, gravitational acceleration measurements support planetary science, as exemplified by NASA's InSight mission (2018-2022), which deployed seismometers on Mars to study its interior, operating under a surface g of 3.71 m/s² confirmed by mission models and orbit-derived data. In technology, micro-g accelerometers—capable of sensing accelerations below 10^{-6} g—enable everyday applications like tilt and orientation detection in smartphones via MEMS devices that measure the gravity vector for auto-rotation and augmented reality features. For spacecraft, these sensors provide precise attitude control by detecting micro-accelerations from thrusters or drag, as in drag-free systems that maintain orientation for missions like GRACE, ensuring stable pointing with errors under 10^{-7} rad/s.

References

  1. [1]
    Weight and Mass
    The weight is a measure of the gravitational force on the object. It is equal to the mass times the gravitational acceleration.
  2. [2]
    standard acceleration of gravity - CODATA Value
    standard acceleration of gravity $g_{\rm n}$ ; Numerical value, 9.806 65 m s ; Standard uncertainty, (exact).
  3. [3]
    Sir Isaac Newton - The Universal Law of Gravitation - Physics
    The acceleration due to gravity is approximately the product of the universal gravitational constant G and the mass of the Earth M, divided by the radius of the ...<|control11|><|separator|>
  4. [4]
  5. [5]
  6. [6]
  7. [7]
    [PDF] Gravity on Earth - National Science Foundation
    Gravity is measured as how fast objects accelerate towards each other. The average gravitational pull of the Earth is 9.8 meters per second squared (m/s2). The ...
  8. [8]
    What is gravity and where is it the strongest in the United States?
    The Earth's average is 9.80 m/s2 (32 ft/s2) which is generally reported as the acceleration of gravity on Earth. This means a dropped object will speed up 32ft/ ...Missing: definition | Show results with:definition
  9. [9]
    Newton's Law of Gravity
    Jan 25, 1999 · Gravitational Acceleration: g = (G × Mass)/(distance from the center)2. Comparing gravitational accelerations: acceleration at position A = ...Missing: definition | Show results with:definition
  10. [10]
    3.5: Free Fall – University Physics Volume 1
    Free fall is the motion of an object falling in a gravitational field, with constant acceleration due to gravity, when air resistance and friction are ...
  11. [11]
    Freely falling objects
    All freely falling objects accelerate at approximately the same rate. This acceleration is denoted by g. Its direction is downwards, towards the center of the ...
  12. [12]
    2.7 Falling Objects – College Physics - UCF Pressbooks
    In free fall, objects fall with constant acceleration due to gravity, independent of mass, if air resistance is negligible. The motion is one-dimensional.
  13. [13]
    Falling Object with Air Resistance | Glenn Research Center - NASA
    Jul 18, 2024 · Gravitational Acceleration​​ The value of g is 9.8 meters per square second on the surface of the earth. The gravitational acceleration decreases ...Missing: definition | Show results with:definition
  14. [14]
    Gravity: Notes: Units Associated with Gravitational Acceleration
    A Gal is defined as a centimeter per second squared. Thus, the Earth's gravitational acceleration is approximately 980 Gals. The Gal is named after Galileo ...
  15. [15]
    Motion of Free Falling Object | Glenn Research Center - NASA
    Jul 3, 2025 · An object that falls through a vacuum is subjected to only one external force, the gravitational force, expressed as the weight of the object.
  16. [16]
    Newton's Law of Gravitation - Neil Gehrels Swift Learning Center
    Dec 11, 2018 · The acceleration due to gravity does not depend on the mass of the object falling, but the force it feels, and thus the object's weight, does.Missing: definition | Show results with:definition
  17. [17]
    13.1 Newton's Law of Universal Gravitation - OpenStax
    Sep 19, 2016 · The constant G is called the universal gravitational constant and Cavendish determined it to be G = 6.67 × 10 −11 N · m 2 /kg 2 G = 6.67 × ...
  18. [18]
    Newtonian constant of gravitation - CODATA Value
    Newtonian constant of gravitation $ G $. Numerical value, 6.674 30 x 10-11 m3 kg-1 s-2. Standard uncertainty, 0.000 15 x 10-11 m3 kg-1 s-2.
  19. [19]
    Newton's Philosophiae Naturalis Principia Mathematica
    Dec 20, 2007 · The historical context in which Newton wrote the Principia ... law of universal gravity from the phenomena of orbital motion in Book 3.
  20. [20]
    [PDF] Geophysical gravity - 1 Earth's orbit in the Solar System
    The International Gravity Formula (1967) – Helmhertz' equation obtains gTz as a function of latitude in accord with Clairault's theorem,. gTz = 9.780327(1.0+ ...
  21. [21]
    Centrifugal acceleration
    The centrifugal acceleration causes the magnitude of the apparent gravitational acceleration on the Earth's surface to vary by about $0.3$ percent.
  22. [22]
    [PDF] adjusted gravity control (grav adju) data - National Geodetic Survey
    The correction for the sea level height of the gravity station (+0.3086 milligals/meter) is called the free-air correction.
  23. [23]
    [PDF] Introduction to Potential Fields: Gravity - USGS.gov
    Many of these effects are caused by known sources, such as the Earth's rota- tion, distance from the Earth's center, topographic relief, and tidal variation.
  24. [24]
    Gravity Anomaly Maps and The Geoid - GRACE Fact Sheet
    Mar 30, 2004 · Gravity anomalies are often due to unusual concentrations of mass in a region. For example, the presence of mountain ranges will usually cause ...
  25. [25]
    Chapter 3 Earth Tides and Tidal Deformations - ScienceDirect.com
    ... gravity owing t o the earth's deformation ranges up to 0.04 mgal. Maximum variations in gravity occurring over a tidal cycle thus are 0.24 + 0.04 = 0.3 mgal.
  26. [26]
    Gravity - Acceleration, Earth, Moon - Britannica
    Sep 29, 2025 · Those are the diurnal and semidiurnal tidal variations. For most purposes it is necessary to know only the variation of gravity with time at ...
  27. [27]
    Constraints on Secular Geocenter Velocity From Absolute Gravity ...
    Dec 23, 2022 · ... effect of the atmosphere based on local air-pressure (1 mbar = 0.3 µGal). An exception to this is the site at Churchill on the coast of ...
  28. [28]
    Reducing hydrological disturbances in absolute gravity observations ...
    Feb 10, 2022 · In this work, we compute hydrological gravity effects from global hydrological models for the far zone and a combination of regional run-off models.
  29. [29]
    [PDF] ACCURACY OF THE RELATIVE GRAVITY MEASUREMENT
    External influences cause real gravity changes and it is necessary to remove them from the measurement. These effects are independent on the used gravimeter.
  30. [30]
    [PDF] A GUIDE TO HIGH PRECISION LAND GRAVIMETER SURVEYS
    All relative, spring-balance gravimeters display long term drift related to the relaxation of spring tensions and the aging of critical components, me-.
  31. [31]
    [PDF] NOAA Technical Memorandum NOS NGS 93 Absolute Gravity ...
    Sep 20, 2023 · field to roughly two parts in 1 billion (or ± 2 µGal, see below). The A10 instrument operates on similar technology but has the advantages ...
  32. [32]
    How zero-gravity parabolic flights work
    Parabolic flights reproduce gravity-free conditions in an aircraft by alternating upward and downward arcs interspersed with level flight.Missing: measurement | Show results with:measurement
  33. [33]
    [PDF] Playing Baseball on the Earth-like Planet Kepler-22b!
    ... gravity on a planetary surface is given by the formula. 2. GM a. R. = where M is in kilograms, R is in meters and G is the Newtonian Constant of Gravity with a.
  34. [34]
    Weight Equation | Glenn Research Center - NASA
    Nov 20, 2023 · The weight W, or gravitational force, is then just the mass of an object times the gravitational acceleration.
  35. [35]
    Planet Compare - Solar System Exploration - NASA
    Surface Gravity(m/s2), 3.7, 8.87, 9.80665, 3.71, 24.79, 10.4*, 8.87, 11.15. Escape Velocity(km/h), 15,300, 37,296, 40,284, 18,108, 216,720, 129,924, 76,968 ...Missing: source | Show results with:source
  36. [36]
    [PDF] Chapter 2 - The Earth's Gravitational field
    The gravitational acceleration was first determined by Galileo; the magnitude of varies over the surface of Earth but a useful ball-park figure is = 9.8 ms-2 ( ...
  37. [37]
    Gravitational Force Equations to Know for Honors Physics - Fiveable
    The formula is ( v_e = \sqrt{\frac{2GM}{r}} ), where ( v_e ) is escape velocity, ( G ) is the gravitational constant, ( M ) is the mass of the planet, and ( r ) ...Missing: source | Show results with:source
  38. [38]
    Neutron stars - UMD Astronomy - University of Maryland
    The surface gravity is about 10^11 times Earth's, and the magnetic field is ... If the magnetic field at the neutron star's surface exceeds about 10^8 ...
  39. [39]
    [PDF] Introductory Lectures on Black Hole Thermodynamics - UMD Physics
    , so a larger black hole has a smaller surface gravity. This happens to be identical to the Newtonian surface gravity of a spherical mass M with radius equal to.
  40. [40]
    By the Numbers | Earth's Moon - Solar System Exploration - NASA
    5.1006 x 108km2. SURFACE GRAVITY. The gravitational acceleration experienced at its surface at the equator. 1.624m/s2. 5.328ft/s2. 1.624m/s2. 9.80665m/s2.
  41. [41]
    Sun Fact Sheet - Radio Jove - NASA
    Jan 16, 1998 · Bulk parameters. Sun Earth Ratio (Sun/Earth) Mass (1024 kg) 1,989,100. ... 0.255 Surface gravity (eq.) (m/s2) 274.0 9.78 28.0 Escape velocity ...
  42. [42]
  43. [43]
    Physics - Satellite Confirms the Principle of Falling
    Sep 14, 2022 · The equivalence principle was first discussed in the early 17th century by Kepler and Galileo, before reappearing later in Newton's theory of ...
  44. [44]
    Tests of the weak equivalence principle - IOPscience
    From the probably apocryphal 16th century demonstrations by Galileo at Pisa's leaning tower to the sensitive torsion-balance measurements of today (both ...
  45. [45]
    The happiest thought of my life.
    In 1907, only two years after the publication of his Special Theory of Relativity, Einstein wrote a paper attempting to modify Newton's theory of gravitation to ...
  46. [46]
    The elevator, the rocket, and gravity: the equivalence principle
    This follows from what Einstein formulated as his equivalence principle which, in turn, is inspired by the consequences of free fall.
  47. [47]
    Atom-Interferometric Test of the Equivalence Principle at the Level
    Nov 2, 2020 · With a resolution of 1.4 × 10 - 11 g per shot and 15 s cycle time, the interferometer attains the highest sensitivity to η of any laboratory ...
  48. [48]
    Equivalence Principle | The Eöt-Wash Group
    The equivalence principle (EP) states that all laws of special relativity hold locally, regardless of the kind of matter involved.<|separator|>
  49. [49]
    General formalism of the quantum equivalence principle - Nature
    Aug 1, 2023 · A consistent theory of quantum gravity will require a fully quantum formulation of the classical equivalence principle.
  50. [50]
    Relativity in the Global Positioning System - PMC - NIH
    GPS satellite clock data for 25 satellites based on ground system observations. Data rate is every 5 minutes, in GPS time; accuracy ranges between 5 and 10 cm.
  51. [51]
    6.3: Galileo's Falling Bodies - Physics LibreTexts
    Oct 31, 2022 · Galileo, and many other scientists and thinkers through the centuries have used ramps to study gravity. This apparatus was originally developed ...
  52. [52]
    June 16, 1657: Christiaan Huygens Patents the First Pendulum Clock
    Jun 16, 2017 · Huygen completed a prototype of his first pendulum clock by the end of 1656, and hired a local clockmaker named Salomon Coster to construct ...
  53. [53]
    Christiaan Huygens (1629 - 1695) - Biography - MacTutor
    In 1656 he patented the first pendulum clock, which greatly increased the accuracy of time measurement. His work on the pendulum was related to other ...
  54. [54]
    How a Scottish mountain weighed the planet - BBC
    Oct 7, 2021 · Crucially, Maskelyne needed to prove that Schiehallion's gravity was drawing the bob of the pendulum away from its vertical position. Maskelyne ...Missing: deflection source<|control11|><|separator|>
  55. [55]
    How Do You Measure the Strength of Gravity? | NIST
    Feb 24, 2025 · One of the earliest known attempts was made in 1798 by English scientist Henry Cavendish, who used a device called a torsion balance to measure ...
  56. [56]
    Kater Reversible Gravity Pendulum
    The Kater pendulum, invented in 1817, is a reversible free swinging pendulum used to measure the local acceleration of gravity.
  57. [57]
    [PDF] Determination of g by Kater pendulum - University of Glasgow
    The reversible, or Kater, pendulum was devised by Henry Kater in 1817 to be used for determining the local value of g. Its advantage lay in the fact that ...
  58. [58]
    Historical development of the gravity method in exploration - Available
    First pendulum clock and its use to measure absolute gravity ...
  59. [59]
    FG5-X Absolute Gravity Meters - Micro g LaCoste
    Featuring: Extended free-fall chamber length; Improved Electronics; Improved reliability; Redesigned drive system. Patented vertical in-line interferometer.
  60. [60]
    A new generation of absolute gravimeters - IOPscience
    We describe the design improvements incorporated in a new generation of absolute gravimeters, the FG5. A vertically oriented (in-line) interferometer design is ...
  61. [61]
    About Us - Micro g LaCoste
    In 2001, LaCoste & Romberg, based in Austin, Texas, a long time producer of metal, spring-based land and air /sea relative gravimeters, and Scintrex merged ...
  62. [62]
    Superconducting Gravimeter - Center for Space Research - The ...
    Superconducting gravimeters are classified as “relative gravimeters” rather the alternative 'absolute gravimeters.” Relative gravimeters are able to constantly ...
  63. [63]
    Satellite Gravimetry: A Review of Its Realization - PMC
    Oct 7, 2021 · In the first forty years of space age, satellite gravimetry was based on the analysis of the orbital motion of satellites.
  64. [64]
    Gravity Method | US EPA
    Dec 12, 2024 · In oil exploration the gravity method is particularly applicable in salt provinces, overthrust and foothills belts, underexplored basins, and ...
  65. [65]
    Microgravity effect of inter-seismic crustal dilatation - Nature
    Nov 5, 2022 · Dilatation strain is defined as a fractional decrease in area (horizontal dilatation), or volume (volumetric dilatation) caused by deformation.
  66. [66]
    A Minimally Cemented Shallow Crust Beneath InSight - Wright - 2022
    Aug 9, 2022 · Model results confirm that the upper 300 m of Mars beneath InSight ... where g, h, and pf represent Mars' gravitational acceleration (3.71 m/s2), ...
  67. [67]
    Does Smartphone Accelerometer Technology Provide Accurate Data?
    Dec 24, 2022 · The accelerometer sensor measures constant (gravity), time varying (vibrations) and quasi static (tilt) acceleration forces, which affect the ...Missing: micro- | Show results with:micro-
  68. [68]
    Attitude determination & control system design for gravity recovery ...
    In this paper we propose the design of attitude determination and control system that could be implemented on-board GRACE-like satellite.