In physics, theories of gravitation postulate mechanisms of interaction governing the movements of bodies with mass. There have been numerous theories of gravitation since ancient times. The first extant sources discussing such theories are found in ancient Greek philosophy. This work was furthered through the Middle Ages by Indian, Islamic, and European scientists, before gaining great strides during the Renaissance and Scientific Revolution—culminating in the formulation of Newton's law of gravity. This was superseded by Albert Einstein's theory of relativity in the early 20th century.

Greek philosopher Aristotle (fl. 4th century BCE) found that objects immersed in a medium tend to fall at speeds proportional to their weight. Vitruvius (fl. 1st century BCE) understood that objects fall based on their specific gravity. In the 6th century CE, Byzantine Alexandrian scholar John Philoponus modified the Aristotelian concept of gravity with the theory of impetus. In the 7th century, Indian astronomer Brahmagupta spoke of gravity as an attractive force. In the 14th century, European philosophers Jean Buridan and Albert of Saxony—who were influenced by certain Islamic scholars[a]—developed the theory of impetus and linked it to the acceleration and mass of objects. Albert also developed a law of proportion regarding the relationship between the speed of an object in free fall and the time elapsed.

Italians of the 16th century found that objects in free fall tend to accelerate equally. In 1632, Galileo Galilei put forth the basic principle of relativity. The existence of the gravitational constant was explored by various researchers from the mid-17th century, helping Isaac Newton formulate his law of universal gravitation. Newton's classical mechanics were superseded in the early 20th century, when Einstein developed the special and general theories of relativity. An elemental force carrier of gravity is hypothesized in quantum gravity approaches such as string theory, in a potentially unified theory of everything.

Antiquity

Classical antiquity

Heraclitus and Leucippus

Heraclitus
A line engraving of Leucippus
Leucippus

The Ionian Greek philosopher Heraclitus (c. 535 – c. 475 BCE) used the word logos ('word') to describe a kind of law which keeps the cosmos in harmony, moving all objects, including the stars, winds, and waves.[3]

The ancient atomist Leucippus (5th-century BCE) proposed the cosmos was created when a large group of atoms came together and swirled as a vortex. The smaller atoms became the celestial bodies of the cosmos. The larger atoms in the center came together as a membrane from which the Earth was formed.[4][5]

Aristotle

Aristotle
Aristotle found that objects immersed in a medium tend to fall at speeds proportional to their weight and inversely proportional to the density of the medium.[6][7][8]

In the 4th century BCE, Greek philosopher Aristotle taught that there is no effect or motion without a cause. The cause of the downward natural motion of heavy bodies, such as the element earth and water, was related to their nature (gravity), which caused them to move downward toward the center of the (geocentric) universe. For this reason Aristotle supported a spherical Earth, since "every portion of earth has weight until it reaches the centre, and the jostling of parts greater and smaller would bring about not a waved surface, but rather compression and convergence of part and part until the centre is reached".[9] On the other hand, light bodies such as the element fire and air, were moved by their nature (levity) upward toward the celestial sphere of the Moon (see sublunary sphere). Astronomical objects near the fixed stars are composed of aether, whose natural motion is circular. Beyond them is the prime mover, the final cause of all motion in the cosmos.[10][11] In his Physics, Aristotle correctly asserted that objects immersed in a medium tend to fall at speeds proportional to their weight and inversely proportional to the density of the medium.[6][8]

Strato of Lampsacus, Epicurus and Aristarchus of Samos

Greek philosopher Strato of Lampsacus (c. 335 – c. 269 BCE) rejected the Aristotelian belief of "natural places" in exchange for a mechanical view in which objects do not gain weight as they fall, instead arguing that the greater impact was due to an increase in speed.[12][13]

Epicurus (c. 341–270 BCE) viewed weight as an inherent property of atoms which influences their movement.[14] These atoms move downward in constant free fall within an infinite vacuum without resistance at equal speed, regardless of their mass. On the other hand, upward motion is due to atomic collisions.[15] Epicureans deviated from older atomist theories like Democritus' (c. 460–c. 370 BCE) by proposing the idea that atoms may randomly deviate from their expected course.[16]

Greek astronomer Aristarchus of Samos (c. 310 – c. 230 BCE) theorized Earth's rotation around its own axis and the orbit of Earth around the Sun in a heliocentric cosmology.[17] Seleucus of Seleucia (c. 190 – c. 150 BCE) supported his cosmology[17] and also described gravitational effects of the Moon on the tidal range.[18]

Archimedes

The 3rd-century-BCE Greek physicist Archimedes (c. 287 – c. 212 BCE) discovered the centre of mass of a triangle.[19] He also postulated that if the centres of gravity of two equal weights was not the same, it would be located in the middle of the line that joins them.[20] In On Floating Bodies, Archimedes claimed that for any object submerged in a fluid there is an equivalent upward buoyant force to the weight of the fluid displaced by the object's volume.[21] The fluids described by Archimedes are not self-gravitating, since he assumes that "any fluid at rest is the surface of a sphere whose centre is the same as that of the Earth".[22][23]

Hipparchus of Nicaea, Lucretius and Vitruvius

Greek astronomer Hipparchus of Nicaea (c.190 – c. 120 BCE) also rejected Aristotelian physics and followed Strato in adopting some form of theory of impetus to explain motion.[24][25] The poem De rerum natura by Lucretius (c. 99 – c. 55 BCE) asserts that more massive bodies fall faster in a medium because the latter resists less, but in a vacuum fall with equal speed.[26] Roman engineer and architect Vitruvius (c. 85 – c. 15 BCE) contends in his De architectura that gravity is not dependent on a substance's weight but rather on its 'nature' (cf. specific gravity):

If the quicksilver is poured into a vessel, and a stone weighing one hundred pounds is laid upon it, the stone swims on the surface, and cannot depress the liquid, nor break through, nor separate it. If we remove the hundred pound weight, and put on a scruple of gold, it will not swim, but will sink to the bottom of its own accord. Hence, it is undeniable that the gravity of a substance depends not on the amount of its weight, but on its nature.[27][28]

Plutarch, Pliny the Elder, and Claudius Ptolemy

Pliny the Elder

Greek philosopher Plutarch (c. 46 – 120 CE) attested the existence of Roman astronomers who rejected Aristotelian physics, "even contemplating theories of inertia and universal gravitation",[29][30] and suggested that gravitational attraction was not unique to the Earth.[31] The gravitational effects of the Moon on the tides were noticed by Pliny the Elder (23–79 CE) in his Naturalis Historia[32] and Claudius Ptolemy (100 – c. 170 CE) in his Tetrabiblos.[33]

Byzantine era

John Philoponus

In the 6th century CE, the Byzantine Alexandrian scholar John Philoponus proposed the theory of impetus, which modifies Aristotle's theory that "continuation of motion depends on continued action of a force" by incorporating a causative force which diminishes over time. In his commentary on Aristotle's Physics that "if one lets fall simultaneously from the same height two bodies differing greatly in weight, one will find that the ratio of the times of their motion does not correspond to the ratios of their weights, but the difference in time is a very small one".[34]

Indian subcontinent

Brahmagupta

Ujjain, Ram Ghat, home to Brahmagupta and Bhaskaracharya

The Indian mathematician/astronomer Brahmagupta (c. 598 – c. 668 CE) first described gravity as an attractive force, using the term "gurutvākarṣaṇam (गुरुत्वाकर्षणम्)" to describe it:[35][36][37][38]

The earth on all its sides is the same; all people on the earth stand upright, and all heavy things fall down to the earth by a law of nature, for it is the nature of the earth to attract and to keep things, as it is the nature of water to flow ... If a thing wants to go deeper down than the earth, let it try. The earth is the only low thing, and seeds always return to it, in whatever direction you may throw them away, and never rise upwards from the earth.[39][40][b]

Bhāskarāchārya

Another famous Indian mathematician and astronomer, Bhāskarā II (Bhāskarāchārya, "Bhāskara, the teacher", c. 1114 – c. 1185), describes gravity as an inherent attractive property of Earth in the section Golādhyāyah (On Spherics) of his treatise Siddhānta Shiromani:

The property of attraction is inherent in the Earth. By this property the Earth attracts any unsupported heavy thing towards it: The thing appears to be falling but it is in a state of being drawn to Earth. ... It is manifest from this that ... people situated at distances of a fourth part of the circumference [of earth] from us or in the opposite hemisphere, cannot by any means fall downwards [in space].[41][42]

Islamic world

Ibn Sina

Ibn Sina

In the 11th century CE, Persian polymath Ibn Sina (Avicenna) agreed with Philoponus' theory that "the moved object acquires an inclination from the mover" as an explanation for projectile motion.[43] Ibn Sina then published his own theory of impetus in The Book of Healing (c. 1020). Unlike Philoponus, who believed that it was a temporary virtue that would decline even in a vacuum, Ibn Sina viewed it as a persistent, requiring external forces such as air resistance to dissipate it.[44][45][1] Ibn Sina made distinction between 'force' and 'inclination' (mayl), and argued that an object gained mayl when the object is in opposition to its natural motion. He concluded that continuation of motion is attributed to the inclination that is transferred to the object, and that object will be in motion until the mayl is spent.[46] The Iraqi polymath Ibn al-Haytham describes gravity as a force in which heavier body moves towards the centre of the earth. He also describes the force of gravity will only move towards the direction of the centre of the earth not in different directions.[47]

Al-Biruni

Al-Biruni

Another 11th-century Persian polymath, Al-Biruni, proposed that heavenly bodies have mass, weight, and gravity, just like the Earth. He criticized both Aristotle and Ibn Sina for holding the view that only the Earth has these properties.[48] The 12th-century scholar Al-Khazini suggested that the gravity an object contains varies depending on its distance from the centre of the universe (referring to the centre of the Earth). Al-Biruni and Al-Khazini studied the theory of the centre of gravity, and generalized and applied it to three-dimensional bodies. Fine experimental methods were also developed for determining the specific gravity or specific weight of objects, based the theory of balances and weighing.[49]

Abu'l-Barakāt al-Baghdādī

In the 12th century, Abu'l-Barakāt al-Baghdādī adopted and modified Ibn Sina's theory on projectile motion. In his Kitab al-Mu'tabar, Abu'l-Barakat stated that the mover imparts a violent inclination (mayl qasri) on the moved and that this diminishes as the moving object distances itself from the mover.[2] According to Shlomo Pines, al-Baghdādī's theory of motion was "the oldest negation of Aristotle's fundamental dynamic law [namely, that a constant force produces a uniform motion], [and is thus an] anticipation in a vague fashion of the fundamental law of classical mechanics [namely, that a force applied continuously produces acceleration]."[50]

European Renaissance

14th century

A 14th century illustration from Gautier de Metz's L'Image du monde showing the gravitational attraction of the Earth at its antipodes.

Jean Buridan, the Oxford Calculators, Albert of Saxony

In the 14th century, both the French philosopher Jean Buridan and the Oxford Calculators (the Merton School) of the Merton College of Oxford rejected the Aristotelian concept of gravity.[51][c] They attributed the motion of objects to an impetus (akin to momentum), which varies according to velocity and mass;[51] Buridan was influenced in this by Ibn Sina's Book of Healing.[1] Buridan and the philosopher Albert of Saxony (c. 1320–1390) adopted Abu'l-Barakat's theory that the acceleration of a falling body is a result of its increasing impetus.[2] Influenced by Buridan, Albert developed a law of proportion regarding the relationship between the speed of an object in free fall and the time elapsed.[52] He also theorized that mountains and valleys are caused by erosion[d]—displacing the Earth's centre of gravity.[53][e]

Uniform and difform motion

The roots of Domingo de Soto's expression uniform difform motion [uniformly accelerated motion] lies in the Oxford Calculators terms "uniform" motion and "difform" motion.[55] "Uniform" motion was used differently then than it would be now. "Uniform" motion might have referred both to constant speed and to motion in which all parts of a body are moving at equal speed. Apparently, the Calculators did not illustrate the different types of motion with real-world examples.[55] John of Holland at the University of Prague, illustrated uniform motion with what would later be called uniform velocity, but also with a falling stone (all parts moving at the same speed), and with a sphere in uniform rotation. He did, however, make distinctions between different kinds of "uniform" motion. Difform motion was exemplified by walking at increasing speed.[55]

Mean speed theorem

Nicole Oresme

Also in the 14th century, the Merton School developed the mean speed theorem; a uniformly accelerated body starting from rest travels the same distance as a body with uniform speed whose speed is half the final velocity of the accelerated body. The mean speed theorem was proved by Nicole Oresme (c. 1323–1382) and would be influential in later gravitational equations.[51] Written as a modern equation:

However, since small time intervals could not be measured, the relationship between time and distance was not so evident as the equation suggests. More generally; equations, which were not widely used until after Galileo's time, imply a clarity that was not there.

15th–17th century

Leonardo da Vinci

Leonardo da Vinci

Leonardo da Vinci (1452–1519) made drawings recording the acceleration of falling objects.[56] He wrote that the "mother and origin of gravity" is energy. He describes two pairs of physical powers which stem from a metaphysical origin and have an effect on everything: abundance of force and motion, and gravity and resistance. He associates gravity with the 'cold' classical elements, water and earth, and calls its energy infinite.[57][f] In Codex Arundel, Leonardo recorded that if a water-pouring vase moves transversally (sideways), simulating the trajectory of a vertically falling object, it produces a right triangle with equal leg length, composed of falling material that forms the hypotenuse and the vase trajectory forming one of the legs.[59] On the hypotenuse, Leonardo noted the equivalence of the two orthogonal motions, one effected by gravity and the other proposed by the experimenter.[59]

Nicolaus Copernicus, Petrus Apianus

Nicolaus Copernicus

By 1514, Nicolaus Copernicus had written an outline of his heliocentric model, in which he stated that Earth's centre is the centre of both its rotation and the orbit of the Moon.[60][g] In 1533, German humanist Petrus Apianus described the exertion of gravity:[h]

Since it is apparent that in the descent [along the arc] there is more impediment acquired, it is clear that gravity is diminished on this account. But because this comes about by reason of the position of heavy bodies, let it be called a positional gravity [i.e. gravitas secundum situm][63]

Francesco Beato and Luca Ghini

Luca Ghini

By 1544, according to Benedetto Varchi, the experiments of at least two Italians, Francesco Beato, a Dominican philosopher at Pisa, and Luca Ghini, a physician and botanist from Bologna, had dispelled the Aristotelian claim that objects fall at speeds proportional to their weight.[64]

Domingo de Soto

Domingo de Soto

In 1551, Domingo de Soto theorized that objects in free fall accelerate uniformly in his book Physicorum Aristotelis quaestiones.[65] This idea was subsequently explored in more detail by Galileo Galilei, who derived his kinematics from the 14th-century Merton College and Jean Buridan,[51] and possibly De Soto as well.[65]

Simon Stevin

Simon Stevin

In 1585, Flemish polymath Simon Stevin performed a demonstration for Jan Cornets de Groot, a local politician in the Dutch city of Delft.[66] Stevin dropped two lead balls from the Nieuwe Kerk in that city. From the sound of the impacts, Stevin deduced that the balls had fallen at the same speed. The result was published in 1586.[67][68]

Let us take (as ... Jan Cornets de Groot ... and I have done) two balls of lead, the one ten times larger and heavier than the other, and drop them together from a height of 30 feet on to a board or something on which they give a perceptible sound. Then it will be found that the lighter will not be ten times longer on its way than the heavier, but that they fall together on to the board so simultaneously that their two sounds seem to be one and the same. ... Therefore Aristotle ... is wrong.

— Simon Stevin, De Beghinselen der Weeghconst

Galileo Galilei

Comparison of the antiquated view and the outcome of the experiment (size of the spheres represent their masses, not their volumes)

Between 1589 and 1592,[69] the Italian scientist Galileo Galilei (then professor of mathematics at the University of Pisa) is said to have dropped "unequal weights of the same material" from the Leaning Tower of Pisa to demonstrate that their time of descent was independent of their mass, according to a biography by Galileo's pupil Vincenzo Viviani, composed in 1654 and published in 1717.[70][71]: 19–21 [72][73] The basic premise had already been demonstrated by Italian experimenters a few decades earlier.

According to the story, Galileo discovered through this experiment that the objects fell with the same acceleration, proving his prediction true, while at the same time disproving Aristotle's theory of gravity (which states that objects fall at speed proportional to their mass). Most historians consider it to have been a thought experiment rather than a physical test.[74]

Galileo successfully applied mathematics to the acceleration of falling objects,[75] correctly hypothesizing in a 1604 letter to Paolo Sarpi that the distance of a falling object is proportional to the square of the time elapsed.[76][i]

I have arrived at a proposition, ... namely, that spaces traversed in natural motion are in the squared proportion of the times.

— Galileo Galilei, Letter to Paolo Sarpi

Written with modern symbols: st2

The result was published in Two New Sciences in 1638. In the same book, Galileo suggested that the slight variance of speed of falling objects of different mass was due to air resistance, and that objects would fall completely uniformly in a vacuum.[77] The relation of the distance of objects in free fall to the square of the time taken was confirmed by Italian Jesuits Grimaldi and Riccioli between 1640 and 1650. They also made a calculation of the gravity of Earth by recording the oscillations of a pendulum.[78]

Johannes Kepler

Johannes Kepler

In his Astronomia nova (1609), Johannes Kepler proposed an attractive force of limited radius between any "kindred" bodies:

Gravity is a mutual corporeal disposition among kindred bodies to unite or join together; thus the earth attracts a stone much more than the stone seeks the earth. (The magnetic faculty is another example of this sort).... If two stones were set near one another in some place in the world outside the sphere of influence of a third kindred body, these stones, like two magnetic bodies, would come together in an intermediate place, each approaching the other by a space proportional to the bulk [moles] of the other....[79]

Evangelista Torricelli

A disciple of Galileo, Evangelista Torricelli reiterated Aristotle's model involving a gravitational centre, adding his view that a system can only be in equilibrium when the common centre itself is unable to fall.[62]

European Enlightenment

The relation of the distance of objects in free fall to the square of the time taken was confirmed by Francesco Maria Grimaldi and Giovanni Battista Riccioli between 1640 and 1650. They also made a calculation of the gravity of Earth constant by recording the oscillations of a pendulum.[80]

Mechanical explanations

In 1644, René Descartes proposed that no empty space can exist and that a continuum of matter causes every motion to be curvilinear. Thus, centrifugal force thrusts relatively light matter away from the central vortices of celestial bodies, lowering density locally and thereby creating centripetal pressure.[81][82] Using aspects of this theory, between 1669 and 1690, Christiaan Huygens designed a mathematical vortex model. In one of his proofs, he shows that the distance elapsed by an object dropped from a spinning wheel will increase proportionally to the square of the wheel's rotation time.[83] In 1671, Robert Hooke speculated that gravitation is the result of bodies emitting waves in the aether.[84][j] Nicolas Fatio de Duillier (1690) and Georges-Louis Le Sage (1748) proposed a corpuscular model using some sort of screening or shadowing mechanism. In 1784, Le Sage posited that gravity could be a result of the collision of atoms, and in the early 19th century, he expanded Daniel Bernoulli's theory of corpuscular pressure to the universe as a whole.[85] A similar model was later created by Hendrik Lorentz (1853–1928), who used electromagnetic radiation instead of corpuscles.

English mathematician Isaac Newton used Descartes' argument that curvilinear motion constrains inertia,[86] and in 1675, argued that aether streams attract all bodies to one another.[k] Newton (1717) and Leonhard Euler (1760) proposed a model in which the aether loses density near mass, leading to a net force acting on bodies.[citation needed] Further mechanical explanations of gravitation (including Le Sage's theory) were created between 1650 and 1900 to explain Newton's theory, but mechanistic models eventually fell out of favor because most of them lead to an unacceptable amount of drag (air resistance), which was not observed. Others violate the energy conservation law and are incompatible with modern thermodynamics.[87]

'Weight' before Newton

Before Newton, 'weight' had the double meaning 'amount' and 'heaviness'.[88]

What we now know as mass was until the time of Newton called “weight.” ... A goldsmith believed that an ounce of gold was a quantity of gold. ... But the ancients believed that a beam balance also measured “heaviness” which they recognized through their muscular senses. ... Mass and its associated downward force were believed to be the same thing. Kepler formed a [distinct] concept of mass (“amount of matter” (copia materiae)), but called it “weight” as did everyone at that time.

— K. M. Browne, The pre-Newtonian meaning of the word “weight”

Mass as distinct from weight

Portrait of Isaac Newton (1642–1727) by Godfrey Kneller (1689)

In 1686, Newton gave the concept of mass its name. In the first paragraph of Principia, Newton defined quantity of matter as “density and bulk conjunctly”, and mass as quantity of matter.[89]

The quantity of matter is the measure of the same, arising from its density and bulk conjunctly. ... It is this quantity that I mean hereafter everywhere under the name of body or mass. And the same is known by the weight of each body; for it is proportional to the weight.

— Isaac Newton, Mathematical principles of natural philosophy, Definition I.

Newton's law of universal gravitation

In 1679, Robert Hooke wrote to Isaac Newton of his hypothesis concerning orbital motion, which partly depends on an inverse-square force.[90][l] In 1684, both Hooke and Newton told Edmond Halley that they had proven the inverse-square law of planetary motion, in January and August, respectively.[92] While Hooke refused to produce his proofs, Newton was prompted to compose De motu corporum in gyrum ('On the motion of bodies in an orbit'), in which he mathematically derives Kepler's laws of planetary motion.[92] In 1687, with Halley's support (and to Hooke's dismay), Newton published Philosophiæ Naturalis Principia Mathematica (Mathematical Principles of Natural Philosophy), which hypothesizes the inverse-square law of universal gravitation.[92] In his own words:

I deduced that the forces which keep the planets in their orbs must be reciprocally as the squares of their distances from the centres about which they revolve; and thereby compared the force requisite to keep the moon in her orb with the force of gravity at the surface of the earth; and found them to answer pretty nearly.

Newton's original formula was:

where the symbol means "is proportional to". To make this into an equal-sided formula or equation, there needed to be a multiplying factor or constant that would give the correct force of gravity no matter the value of the masses or distance between them – the gravitational constant. Newton would need an accurate measure of this constant to prove his inverse-square law. Reasonably accurate measurements were not available in until the Cavendish experiment by Henry Cavendish in 1797.[93]

In Newton's theory[94] (rewritten using more modern mathematics) the density of mass generates a scalar field, the gravitational potential in joules per kilogram, by

Using the Nabla operator for the gradient and divergence (partial derivatives), this can be conveniently written as:

This scalar field governs the motion of a free-falling particle by:

At distance r from an isolated mass M, the scalar field is

The Principia sold out quickly, inspiring Newton to publish a second edition in 1713.[95][96] However the theory of gravity itself was not accepted quickly.

The theory of gravity faced two barriers. First scientists like Gottfried Wilhelm Leibniz complained that it relied on action at a distance, that the mechanism of gravity was "invisible, intangible, and not mechanical".[97]: 339 [98]: 144  The French philosopher Voltaire countered these concerns, ultimately writing his own book to explain aspects of it to French readers in 1738, which helped to popularize Newton's theory.[99]

Second, detailed comparisons with astronomical data were not initially favorable. Among the most conspicuous issue was the so-called great inequality of Jupiter and Saturn. Comparisons of ancient astronomical observations to those of the early 1700s implied that the orbit of Saturn was increasing in diameter while that of Jupiter was decreasing. Ultimately this meant Saturn would exit the Solar System and Jupiter would collide with other planets or the Sun. The problem was tackled first by Leonhard Euler in 1748, then Joseph-Louis Lagrange in 1763, by Pierre-Simon Laplace in 1773. Each effort improved the mathematical treatment until the issue was resolved by Laplace in 1784 approximately 100 years after Newton's first publication on gravity. Laplace showed that the changes were periodic but with immensely long periods beyond any existing measurements.[100]: 144 

Successes such the solution to the great inequality of Jupiter and Saturn mystery accumulated. In 1755, Prussian philosopher Immanuel Kant published a cosmological manuscript based on Newtonian principles, in which he develops an early version of the nebular hypothesis.[101] Edmond Halley proposed that similar looking objects appearing every 76 years was in fact a single comet. The appearance of the comet in 1759, now named after him, within a month of predictions based on Newton's gravity greatly improved scientific opinion of the theory.[102] Newton's theory enjoyed its greatest success when it was used to predict the existence of Neptune based on motions of Uranus that could not be accounted by the actions of the other planets. Calculations by John Couch Adams and Urbain Le Verrier both predicted the general position of the planet. In 1846, Le Verrier sent his position to Johann Gottfried Galle, asking him to verify it. The same night, Galle spotted Neptune near the position Le Verrier had predicted.[103]

Not every comparison was successful. By the end of the 19th century, Le Verrier showed that the orbit of Mercury could not be accounted for entirely under Newtonian gravity, and all searches for another perturbing body (such as a planet orbiting the Sun even closer than Mercury) were fruitless.[104] Even so, Newton's theory is thought to be exceptionally accurate in the limit of weak gravitational fields and low speeds.

At the end of the 19th century, many tried to combine Newton's force law with the established laws of electrodynamics (like those of Wilhelm Eduard Weber, Carl Friedrich Gauss, and Bernhard Riemann) in order to explain the anomalous perihelion precession of Mercury. In 1890, Maurice Lévy succeeded in doing so by combining the laws of Weber and Riemann, whereby the speed of gravity is equal to the speed of light. In another attempt, Paul Gerber (1898) succeeded in deriving the correct formula for the perihelion shift (which was identical to the formula later used by Albert Einstein). These hypotheses were rejected because of the outdated laws they were based on, being superseded by those of James Clerk Maxwell.[87]

Modern era

In 1900, Hendrik Lorentz tried to explain gravity on the basis of his ether theory and Maxwell's equations. He assumed, like Ottaviano Fabrizio Mossotti and Johann Karl Friedrich Zöllner, that the attraction of opposite charged particles is stronger than the repulsion of equal charged particles. The resulting net force is exactly what is known as universal gravitation, in which the speed of gravity is that of light. Lorentz calculated that the value for the perihelion advance of Mercury was much too low.[105]

In the late 19th century, Lord Kelvin pondered the possibility of a theory of everything.[106] He proposed that every body pulsates, which might be an explanation of gravitation and electric charges. His ideas were largely mechanistic and required the existence of the aether, which the Michelson–Morley experiment failed to detect in 1887. This, combined with Mach's principle, led to gravitational models which feature action at a distance.

Albert Einstein developed his revolutionary theory of relativity in papers published in 1905 and 1915; these account for the perihelion precession of Mercury.[104] In 1914, Gunnar Nordström attempted to unify gravity and electromagnetism in his theory of five-dimensional gravitation.[m] General relativity was proven in 1919, when Arthur Eddington observed gravitational lensing around a solar eclipse, matching Einstein's equations. This resulted in Einstein's theory superseding Newtonian physics.[107] Thereafter, German mathematician Theodor Kaluza promoted the idea of general relativity with a fifth dimension, which in 1921 Swedish physicist Oskar Klein gave a physical interpretation of in a prototypical string theory, a possible model of quantum gravity and potential theory of everything.

Albert Einstein in 1921

Einstein's field equations include a cosmological constant to account for the alleged staticity of the universe. However, Edwin Hubble observed in 1929 that the universe appears to be expanding. By the 1930s, Paul Dirac developed the hypothesis that gravitation should slowly and steadily decrease over the course of the history of the universe.[108] Alan Guth and Alexei Starobinsky proposed in 1980 that cosmic inflation in the very early universe could have been driven by a negative pressure field, a concept later coined 'dark energy'—found in 2013 to have composed around 68.3% of the early universe.[109]

In 1922, Jacobus Kapteyn proposed the existence of dark matter, an unseen force that moves stars in galaxies at higher velocities than gravity alone accounts for. It was found in 2013 to have comprised 26.8% of the early universe.[109] Along with dark energy, dark matter is an outlier in Einstein's relativity, and an explanation for its apparent effects is a requirement for a successful theory of everything.

In 1957, Hermann Bondi proposed that negative gravitational mass (combined with negative inertial mass) would comply with the strong equivalence principle of general relativity and Newton's laws of motion. Bondi's proof yielded singularity-free solutions for the relativity equations.[110]

Early theories of gravity attempted to explain planetary orbits (Newton) and more complicated orbits (e.g. Lagrange). Then came unsuccessful attempts to combine gravity and either wave or corpuscular theories of gravity. The whole landscape of physics was changed with the discovery of Lorentz transformations, and this led to attempts to reconcile it with gravity. At the same time, experimental physicists started testing the foundations of gravity and relativity—Lorentz invariance, the gravitational deflection of light, the Eötvös experiment. These considerations led to and past the development of general relativity.

Einstein (1905, 1908, 1912)

In 1905, Albert Einstein published a series of papers in which he established the special theory of relativity and the fact that mass and energy are equivalent. In 1907, in what he described as "the happiest thought of my life", Einstein realized that someone who is in free fall experiences no gravitational field. In other words, gravitation is exactly equivalent to acceleration.

Einstein's two-part publication in 1912[111][112] (and before in 1908) is really only important for historical reasons. By then he knew of the gravitational redshift and the deflection of light. He had realized that Lorentz transformations are not generally applicable, but retained them. The theory states that the speed of light is constant in free space but varies in the presence of matter. The theory was only expected to hold when the source of the gravitational field is stationary. It includes the principle of least action:

where is the Minkowski metric, and there is a summation from 1 to 4 over indices and .

Einstein and Grossmann[113] includes Riemannian geometry and tensor calculus.

The equations of electrodynamics exactly match those of general relativity. The equation

is not in general relativity. It expresses the stress–energy tensor as a function of the matter density.

Lorentz-invariant models (1905–1910)

Based on the principle of relativity, Henri Poincaré (1905, 1906), Hermann Minkowski (1908), and Arnold Sommerfeld (1910) tried to modify Newton's theory and to establish a Lorentz invariant gravitational law, in which the speed of gravity is that of light. As in Lorentz's model, the value for the perihelion advance of Mercury was much too low.[114]

Abraham (1912)

Meanwhile, Max Abraham developed an alternative model of gravity in which the speed of light depends on the gravitational field strength and so is variable almost everywhere. Abraham's 1914 review of gravitation models is said to be excellent, but his own model was poor.

Nordström (1912)

The first approach of Nordström (1912)[115] was to retain the Minkowski metric and a constant value of but to let mass depend on the gravitational field strength . Allowing this field strength to satisfy

where is rest mass energy and is the d'Alembertian,

where is the mass when gravitational potential vanishes and,

where is the four-velocity and the dot is a differential with respect to time.

The second approach of Nordström (1913)[116] is remembered as the first logically consistent relativistic field theory of gravitation ever formulated. (notation from Pais[117] not Nordström):

where is a scalar field,

This theory is Lorentz invariant, satisfies the conservation laws, correctly reduces to the Newtonian limit and satisfies the weak equivalence principle.

Einstein and Fokker (1914)

This theory[118] is Einstein's first treatment of gravitation in which general covariance is strictly obeyed. Writing:

they relate Einstein–Grossmann[113] to Nordström.[116] They also state:

That is, the trace of the stress energy tensor is proportional to the curvature of space.

Between 1911 and 1915, Einstein developed the idea that gravitation is equivalent to acceleration, initially stated as the equivalence principle, into his general theory of relativity, which fuses the three dimensions of space and the one dimension of time into the four-dimensional fabric of spacetime. However, it does not unify gravity with quanta—individual particles of energy, which Einstein himself had postulated the existence of in 1905.

General relativity

Illustration explaining the relevance of the total solar eclipse of 29 May 1919, from the 22 November 1919 edition of The Illustrated London News

In general relativity, the effects of gravitation are ascribed to spacetime curvature instead of to a force. The starting point for general relativity is the equivalence principle, which equates free fall with inertial motion. The issue that this creates is that free-falling objects can accelerate with respect to each other. To deal with this difficulty, Einstein proposed that spacetime is curved by matter, and that free-falling objects are moving along locally straight paths in curved spacetime. More specifically, Einstein and David Hilbert discovered the field equations of general relativity, which relate the presence of matter and the curvature of spacetime. These field equations are a set of 10 simultaneous, non-linear, differential equations. The solutions of the field equations are the components of the metric tensor of spacetime, which describes its geometry. The geodesic paths of spacetime are calculated from the metric tensor.

Notable solutions of the Einstein field equations include:

General relativity has enjoyed much success because its predictions (not called for by older theories of gravity) have been regularly confirmed. For example:

It is believed that neutron star mergers (since detected in 2017)[120] and black hole formation may also create detectable amounts of gravitational radiation.

Quantum gravity

Several decades after the discovery of general relativity, it was realized that it cannot be the complete theory of gravity because it is incompatible with quantum mechanics.[121] Later it was understood that it is possible to describe gravity in the framework of quantum field theory like the other fundamental forces. In this framework, the attractive force of gravity arises due to exchange of virtual gravitons, in the same way as the electromagnetic force arises from exchange of virtual photons.[122][123] This reproduces general relativity in the classical limit, but only at the linearized level and postulating that the conditions for the applicability of Ehrenfest theorem holds, which is not always the case. Moreover, this approach fails at short distances of the order of the Planck length.[121]

See also

References

Footnotes

  1. ^ Ibn Sina and Abu'l-Barakat, respectively[1][2]
  2. ^ The source of this quote is Al-Biruni's India (c. 1030).[39]
  3. ^ This was interpreted as deriving the weight of objects from the pressure of the air below them.[51]
  4. ^ Leonardo da Vinci tested this theory by observing trace fossils,[53] which he used to argue against the myth of a universal flood.[54]
  5. ^ Furthermore, he hypothesized that the planet is in equilibrium when its centre of gravity coincides with that of its mass.[53]
  6. ^ Leonardo did not publish his manuscripts and they had no direct influence on subsequent science.[58]
  7. ^ He accounted for these movements by explaining, "Rotation is natural to a sphere, and by that very act is its shape expressed."[61]
  8. ^ Physicist Pierre Duhem erroneously attributes this to Jordanus Nemorarius, whom he calls the "precursor of Leonardo". Leonardo alludes to Jordanus in his notebooks, but not to any of his theories.[62]
  9. ^ The distance traversed in successive equal intervals of time is calculated with a triangular model whose width (representing maximum velocity) increases by two for every equal section of height (representing time elapsed). This is in part anticipated by the Merton rule.[76]
  10. ^ James Challis repeated this assumption in 1869.
  11. ^ Bernhard Riemann made a similar argument in 1853.
  12. ^ Newton was almost certainly influenced by this correspondence to do his subsequent work on gravitation,[90] although he denied that Hooke had told him of the inverse-square force.[91]
  13. ^ In string theory, dimensions exceeding four allow for the existence of parallel realities—which along with the anthropic principle, help to explain the statistical near-impossibility of our fine-tuned universe.

Citations

  1. ^ a b c Sayili, Aydin (1987). "Ibn Sīnā and Buridan on the Motion of the Projectile". Annals of the New York Academy of Sciences. 500 (1): 477–482. Bibcode:1987NYASA.500..477S. doi:10.1111/j.1749-6632.1987.tb37219.x. S2CID 84784804.
  2. ^ a b c Gutman, Oliver (2003). Pseudo-Avicenna, Liber Celi Et Mundi: A Critical Edition. Brill Publishers. p. 193. ISBN 90-04-13228-7.
  3. ^ Smith, Homer W. (1952). Man and His Gods. New York: Grosset & Dunlap. p. 144.
  4. ^ Furley, David (1987). The Greek Cosmologists: Volume 1, The Formation of the Atomic Theory and its Earliest Critics. Cambridge University Press. pp. 140–141. doi:10.1017/CBO9780511552540. ISBN 0-521-33328-8.
  5. ^ McKirahan, Richard D. (2011) [1994]. Philosophy Before Socrates (2nd ed.). Hackett Publishing Company. pp. 411–412. ISBN 978-1-60384-182-5.
  6. ^ a b "Part I – B. Aristotle's Theory of Free-Fall | Relativity of Gravity". Retrieved 9 June 2023.
  7. ^ Drabkin, Israel E. (1938). "Notes on the Laws of Motion in Aristotle". The American Journal of Philology. 59 (1): 60–84. doi:10.2307/290584. JSTOR 90584.
  8. ^ a b Rovelli, Carlo (2015). "Aristotle's Physics: A Physicist's Look". Journal of the American Philosophical Association. 1 (1): 23–40. arXiv:1312.4057. doi:10.1017/apa.2014.11. ISSN 2053-4477. S2CID 44193681. Retrieved 8 November 2018.
  9. ^ "On the Heavens by Aristotle, Book 2, Part 14". classics.mit.edu. The Internet Classics Archive. Retrieved 23 August 2023.
  10. ^ Grant, Edward (1996). The foundations of modern science in the Middle Ages : their religious, institutional, and intellectual contexts. Internet Archive. Cambridge, England; New York, New York: Cambridge University Press. pp. 60–61. ISBN 978-0-521-56137-2.
  11. ^ Pedersen, Olaf (1993). Early physics and astronomy: a historical introduction. Internet Archive. Cambridge, England; New York, New York: Cambridge University Press. p. 130. ISBN 978-0-521-40340-5.
  12. ^ Carrier, Richard (1 December 2017). The Scientist in the Early Roman Empire. United States and Canada: Pitchstone Publishing. p. 333. ISBN 978-1-63431-107-6. For example, in his lost books On Lightness and Heaviness and On Motion, Strato abandoned the doctrine of 'natural places' in exchange for a more mechanical view of why some objects rise and others fall
  13. ^ Fortenbaugh, William (2017). Strato of Lampsacus: Text, Translation and Discussion. Routledge. ISBN 978-1-351-48792-4. If someone drops a rock [from] a finger's height above the ground, it certainly won't make a visible impact on the ground, but if someone drops it holding it a hundred feet or more, it will have a strong impact. And there is no other reason for that impact. Because it does not have greater weight, nor is it impelled by greater force; but it moves faster.
  14. ^ "Weight in Greek Atomism". Philosophia. 45: 85. 2015.
  15. ^ Laertius, Diogenes. "Letter of Epicurus to Herodotus, (61)". www.attalus.org. Retrieved 26 February 2024.
  16. ^ Berryman, Sylvia (2022), Zalta, Edward N.; Nodelman, Uri (eds.), "Ancient Atomism", The Stanford Encyclopedia of Philosophy (Winter 2022 ed.), Metaphysics Research Lab, Stanford University, retrieved 11 February 2024
  17. ^ a b "Plutarch, Platonicae quaestiones, Question VIII, section 1". www.perseus.tufts.edu. Retrieved 27 August 2023.
  18. ^ Strabo. "Geography — III, 5, 9". penelope.uchicago.edu. Retrieved 27 August 2023.
  19. ^ Neitz, Reviel; Noel, William (13 October 2011). The Archimedes Codex: Revealing The Secrets Of The World's Greatest Palimpsest. United Kingdom: Hachette UK. ISBN 9781780221984.
  20. ^ Tuplin, C. J.; Wolpert, Lewis (2002). Science and Mathematics in Ancient Greek Culture. United Kingdom: Hachette UK. p. xi. ISBN 9780198152484.
  21. ^ "The works of Archimedes". Cambridge, University Press. 1897. p. 257. Retrieved 11 March 2010. Any solid lighter than a fluid will, if placed in the fluid, be so far immersed that the weight of the solid will be equal to the weight of the fluid displaced.
  22. ^ The works of Archimedes. Translated by Heath, T. L. Cambridge: Cambridge University Press. 1897. p. 254. Retrieved 13 November 2017.
  23. ^ Ceccarelli, Marco (17 August 2007). Distinguished Figures in Mechanism and Machine Science: Their Contributions and Legacies. Springer Science & Business Media. p. 13. ISBN 978-1-4020-6366-4.
  24. ^ Sorabji, Richard, ed. (2014). Simplicius: On Aristotle On the Heavens 1.5-9. Translated by Hankinson, R. J. Bloomsbury Publishing. p. 87. ISBN 978-1-4725-0111-0.
  25. ^ Carrier, Richard (1 December 2017). The Scientist in the Early Roman Empire. Pitchstone Publishing (US&CA). ISBN 978-1-63431-107-6. Hipparchus rejected the Aristotlian physics of motion and followed Strato in embracing an early impetus theory
  26. ^ Leonard, William Ellery (ed.). "Lucretius, De Rerum Natura, BOOK II, line 216". Perseus Digital Library. Retrieved 20 August 2023 – via Tufts University.
  27. ^ Vitruvius, Marcus Pollio (1914). "VII". In Howard, Alfred A. (ed.). De Architectura libri decem [Ten Books on Architecture]. Herbert Langford Warren, Nelson Robinson (illus), Morris Hicky Morgan. Harvard University, Cambridge, Massachusetts: Harvard University Press. p. 215.
  28. ^ For another English translation see:The architecture of M. Vitruvius Pollio: translated from the original Latin by W. Newton. Vol. 2. 1791. p. 168.
  29. ^ Carrier, Richard (1 December 2017). The Scientist in the Early Roman Empire. United States and Canada: Pitchstone Publishing. ISBN 978-1-63431-107-6. Plutarch also attests to the existence of Roman philosophers and astronomers who rejected Aristotelian dynamics and were engaging sophisticated debates on the subject, even contemplating theories of inertia and universal gravitation
  30. ^ Taub, Liba Chaia (2008). Aetna and the Moon: Explaining Nature in Ancient Greece and Rome. Oregon State University Press. ISBN 978-0-87071-196-1.
  31. ^ Bakker, Frederik; Palmerino, Carla Rita (1 June 2020). "Motion to the Center or Motion to the Whole? Plutarch's Views on Gravity and Their Influence on Galileo". Isis. 111 (2): 217–238. doi:10.1086/709138. hdl:2066/219256. ISSN 0021-1753. S2CID 219925047.
  32. ^ Pliny the Elder (1893). The Natural History of Pliny. H. G. Bohn. p. 128. ISBN 978-0-598-91073-8.
  33. ^ Ptolemy (1940). "2". Tetrabiblos. Vol. 1. Translated by Robbins, Frank E. Cambridge, Massachusetts: Harvard University Press.
  34. ^ "John Philoponus". www.eoht.info. Retrieved 9 June 2023.
  35. ^ Pickover, Clifford (2008). Archimedes to Hawking: Laws of Science and the Great Minds Behind Them. Oxford University Press. p. 105. ISBN 978-0-19-979268-9.
  36. ^ Bose, Mainak Kumar (1988). Late classical India. A. Mukherjee & Company.[page needed]
  37. ^ Sen, Amartya (2005). The Argumentative Indian. Allen Lane. p. 29. ISBN 978-0-7139-9687-6.
  38. ^ Thurston, Hugh (1993). Early Astronomy. New York: Springer-Verlag. ISBN 978-0-387-94107-3.[page needed]
  39. ^ a b Alberuni's India. London: Kegan Paul, Trench, Trübner & Co., 1910. Electronic reproduction. Vol. 1 and 2. New York: Columbia University Libraries, 2006. p. 272. Retrieved 3 June 2014.
  40. ^ Kitāb al-Jawharatayn al-'atīqatayn al-mā'i'atayn min al-ṣafrā' wa-al-bayḍā': al-dhahab wa-al-fiḍḍah كتاب الجوهرتين العتيقتين المائعتين من الصفراء والبيضاء : الذهب والفضة. Cairo: Maṭba'at Dār al-Kutub wa-al-Wathā'iq al-Qawmīyah bi-al-Qāhirah. 2004. pp. 43–44, 87. OCLC 607846741.
  41. ^ Áryabhat́t́a; Bháskarácárya (1150) [505 & 1150]. "Chapter III ─ Called Bhuvana-kośa or Cosmograghy". Súrya Siddhánta and Siddhánta Shiromańi (in Sanskrit). Translated by Deva Sástri, Bápú; Wilkinson, Lancelot. Calcutta, India: C. B. Lewis, Baptist Mission Press (published 1860). p. 113.
  42. ^ Bháskarácárya (1150). "ভুবনকোষ". Siddhánta Shiromańi: Goládhyáyah (PDF) (in Sanskrit). Calcutta, India.
  43. ^ McGinnis, Jon; Reisman, David C. (2007). Classical Arabic philosophy: an anthology of sources. Hackett Publishing. p. 174. ISBN 978-0-87220-871-1. Retrieved 16 June 2010.
  44. ^ Espinoza, Fernando (2005). "An analysis of the historical development of ideas about motion and its implications for teaching". Physics Education. 40 (2): 141. Bibcode:2005PhyEd..40..139E. doi:10.1088/0031-9120/40/2/002. S2CID 250809354.
  45. ^ Seyyed Hossein Nasr; Mehdi Amin Razavi (1996). The Islamic intellectual tradition in Persia. Routledge. p. 72. ISBN 978-0-7007-0314-2.
  46. ^ Espinoza, Fernando. "An Analysis of the Historical Development of Ideas About Motion and its Implications for Teaching". Physics Education. Vol. 40 (2).
  47. ^ Clagett, Marshall (1961). The Science of Mechanics in the Middle Ages. Vol. 1. Madison, Wisconsin; London, England: University of Wisconsin Press/Oxford University Press. p. 58 – via Internet Archive.
  48. ^ Starr, S. Frederick (2015). Lost Enlightenment: Central Asia's Golden Age from the Arab Conquest to Tamerlane. Princeton University Press. p. 260. ISBN 9780691165851.
  49. ^ Rozhanskaya, Mariam; Levinova, I. S. (1996). "Statics". In Rushdī, Rāshid (ed.). Encyclopedia of the History of Arabic Science. Vol. 2. Psychology Press. pp. 614–642. ISBN 9780415124119. Using a whole body of mathematical methods (not only those inherited from the antique theory of ratios and infinitesimal techniques, but also the methods of the contemporary algebra and fine calculation techniques), Muslim scientists raised statics to a new, higher level. The classical results of Archimedes in the theory of the centre of gravity were generalized and applied to three-dimensional bodies, the theory of ponderable lever was founded and the 'science of gravity' was created and later further developed in medieval Europe. The phenomena of statics were studied by using the dynamic approach so that two trends – statics and dynamics – turned out to be inter-related within a single science, mechanics. The combination of the dynamic approach with Archimedean hydrostatics gave birth to a direction in science which may be called medieval hydrodynamics. ... Numerous fine experimental methods were developed for determining the specific weight, which were based, in particular, on the theory of balances and weighing. The classical works of al-Biruni and al-Khazini can by right be considered as the beginning of the application of experimental methods in medieval science.
  50. ^ Pines, Shlomo (1970). "Abu'l-Barakāt al-Baghdādī, Hibat Allah". Dictionary of Scientific Biography. Vol. 1. New York: Charles Scribner's Sons. pp. 26–28. ISBN 0-684-10114-9.
    (cf. Abel B. Franco (October 2003). "Avempace, Projectile Motion, and Impetus Theory", Journal of the History of Ideas 64 (4), pp. 521–546 [528].)
  51. ^ a b c d e Gillispie 1960, p. 41.
  52. ^ Drake, Stillman (1975). "Free fall from Albert of Saxony to Honoré Fabri". Studies in History and Philosophy of Science Part A. 5 (4): 347–366. Bibcode:1975SHPSA...5..347D. doi:10.1016/0039-3681(75)90007-2. ISSN 0039-3681 – via Academia.edu.
  53. ^ a b c Knight, Kevin (2017). "Albert of Saxony". New Advent. Retrieved 10 July 2019.
  54. ^ Da Vinci, Leonardo (1971). Taylor, Pamela (ed.). The Notebooks of Leonardo da Vinci. New American Library. pp. 136–38, 142–48.
  55. ^ a b c Wallace 2004a, p. 386.
  56. ^ Ouellette, Jennifer (10 February 2023). "Leonardo noted link between gravity and acceleration centuries before Einstein". Ars Technica. Retrieved 11 February 2023.
  57. ^ Da Vinci, Leonardo (1971). Taylor, Pamela (ed.). The Notebooks of Leonardo da Vinci. New American Library. p. 124. Force arises from dearth or abundance; it is the child of physical motion, and the grandchild of spiritual motion, and the mother and origin of gravity. Gravity is limited to the elements of water and earth; but his force is unlimited, and by it infinite worlds might be moved if instruments could be made by which the force be generated.
    Force, with physical motion, and gravity, with resistance, are the four external powers on which all actions of mortals depend.
  58. ^ Capra, Fritjof (2007). The Science of Leonardo. US: Doubleday. pp. 5–6. ISBN 978-0-385-51390-6.
  59. ^ a b Morteza Gharib; Chris Roh; Flavio Noca (1 February 2023). "Leonardo da Vinci's Visualization of Gravity as a Form of Acceleration". Leonardo. 56: 21–27. doi:10.1162/leon_a_02322. S2CID 254299572. Retrieved 16 February 2023.
  60. ^ Durant, Will (2011) [1957]. The Story of Civilization: Volume VI – The Reformation. Simon and Schuster. p. 823. ISBN 9781451647631.
  61. ^ Gillispie 1960, p. 27.
  62. ^ a b Ginzburg, Benjamin (September 1936). "Duhem and Jordanus Nemorarius". Isis. 25 (2). The University of Chicago Press: 341–362. doi:10.1086/347085. JSTOR 225373. S2CID 145152521.
  63. ^ Duhem, Pierre (2012). The Origins of Statics: The Sources of Physical Theory Volume 1. Translated by Leneaux, G. F.; Vagliente, V. N.; Wagener, G. H. Springer Science & Business Media. p. xxiv. ISBN 9789401137300.
  64. ^ Wallace 2004b, p. 121.
  65. ^ a b Wallace, William A. (2018) [2004]. Domingo de Soto and the Early Galileo: Essays on Intellectual History. Abingdon, UK: Routledge. pp. 119, 121–22. ISBN 978-1-351-15959-3.
  66. ^ Icke, V. (2014). Gravity does not exist: A puzzle for the 21st century. Amsterdam University Press. p. 9. Bibcode:2014gdne.book.....I.
  67. ^ Drake, S (1978). Galileo at work: His scientific biography. University of Chicago Press. p. 20. ISBN 9780226162263.
  68. ^ Stevin, S. (1955) [1586]. Dijksterhuis, E. J. (ed.). The Principal Works of Simon Stevin (PDF) (in Dutch and English). Vol. 1. C. V. Swets & Zeitlinger. pp. 509, 511.
  69. ^ Some contemporary sources speculate about the exact date; e.g. Rachel Hilliam gives 1591 (Galileo Galilei: Father of Modern Science, The Rosen Publishing Group, 2005, p. 101).
  70. ^ Vincenzo Viviani (1717), Racconto istorico della vita di Galileo Galilei, p. 606: [...dimostrando ciò con replicate esperienze, fatte dall'altezza del Campanile di Pisa con l'intervento delli altri lettori e filosofi e di tutta la scolaresca... [...Galileo showed this [all bodies, whatever their weights, fall with equal speeds] by repeated experiments made from the height of the Leaning Tower of Pisa in the presence of other professors and all the students...].
  71. ^ Drake, Stillman (2003). Galileo at Work: His Scientific Biography (Facsim. ed.). Mineola (N.Y.): Dover publ. ISBN 9780486495422.
  72. ^ "Sci Tech : Science history: setting the record straight". The Hindu. 30 June 2005. Archived from the original on 2 November 2005. Retrieved 5 May 2009.
  73. ^ Vincenzo Viviani on museo galileo
  74. ^ "El experimento más famoso de Galileo probablemente nunca tuvo lugar". The Conversation. 16 May 2019. Retrieved 17 May 2019.
  75. ^ Gillispie 1960, p. 42.
  76. ^ a b Gillispie 1960, pp. 3–6.
  77. ^ Galilei, Galileo (2015). Dialogues Concerning Two New Sciences. Translated by Crew, Henry. Eastford, CT: Martino Fine Books. p. 72. ISBN 978-1614277941.
  78. ^ J.L. Heilbron, Electricity in the 17th and 18th Centuries: A Study of Early Modern Physics (Berkeley: the University of California Press, 1979), 180.
  79. ^ Kepler, Johannes; William H. Donahue (2004). Selections from Kepler's Astronomia Nova. Santa Fe: Green Lion Press. p. 1. ISBN 1-888009-28-4.
  80. ^ J.L. Heilbron, Electricity in the 17th and 18th Centuries: A Study of Early Modern Physics (Berkeley: University of California Press, 1979), 180.
  81. ^ Gillispie 1960, p. 93.
  82. ^ Descartes, René (1644). Principles of Philosophy.
  83. ^ Gillispie 1960, p. 121.
  84. ^ Taylor, William Bower (1876). "Kinetic Theories of Gravitation" . Smithsonian Report: 205–282.
  85. ^ Gillispie 1960, p. 480.
  86. ^ Gillispie 1960, p. 120.
  87. ^ a b Zenneck, J. (1903). "Gravitation". Encyklopädie der Mathematischen Wissenschaften mit Einschluss ihrer Anwendungen (in German). Vol. 5. Leipzig. pp. 25–67. doi:10.1007/978-3-663-16016-8_2. ISBN 978-3-663-15445-7.{{cite book}}: CS1 maint: location missing publisher (link)
  88. ^ Browne, K. M. (2018). "The pre-Newtonian meaning of the word "weight"; a comment on "Kepler and the origins of pre-Newtonian mass" [Am. J. Phys. 85, 115–123 (2017)]". American Journal of Physics. 86 (6): 471–74. Bibcode:2018AmJPh..86..471B. doi:10.1119/1.5027490. S2CID 125953814.
  89. ^ Newton, I. (1729) [Original work published 1686]. The mathematical principles of natural philosophy. Translated by Motte, A. Printed for Benjamin Motte. pp. 1–2.
  90. ^ a b Cohen, I. Bernard; Smith, George Edwin (2002). The Cambridge Companion to Newton. Cambridge University Press. pp. 11–12, 96–97. ISBN 978-0-521-65696-2.
  91. ^ H. W. Turnbull (ed.), Correspondence of Isaac Newton, Vol. 2 (1676–1687), (Cambridge University Press, 1960), pp. 297–314, 431–448.
  92. ^ a b c Sagan, Carl & Druyan, Ann (1997). Comet. New York: Random House. pp. 52–58. ISBN 978-0-3078-0105-0.
  93. ^ Poynting 1894
  94. ^ Newton, I. (1686). Philosophiæ Naturalis Principia Mathematica (in Latin).
  95. ^ The Correspondence of Isaac Newton, vol. 4, Cambridge University Press 1967, at pp. 519, n.2.
  96. ^ Westfall, Richard S. (1971), Force in Newton's Physics: The Science of Dynamics in the Seventeenth Century. New York: American Elsevier, p. 750.
  97. ^ Hesse, Mary B. (1955). "Action at a Distance in Classical Physics". Isis. 46 (4): 337–353. doi:10.1086/348429. ISSN 0021-1753. JSTOR 227576. S2CID 121166354.
  98. ^ Gillispie, Charles Coulston. The edge of objectivity: An essay in the history of scientific ideas. Princeton University Press, 2016.
  99. ^ Shank, J. B. (2009). "Voltaire". Stanford Encyclopedia of Philosophy.
  100. ^ Whittaker, Edmund T. (1989). A history of the theories of aether & electricity. 2: The Modern Theories 1900-1926 (Repr ed.). New York: Dover Publ. ISBN 978-0-486-26126-3.
  101. ^ Woolfson, M.M. (1993). "Solar System – its origin and evolution". Q. J. R. Astron. Soc. 34: 1–20. Bibcode:1993QJRAS..34....1W.
  102. ^ Hughes, D. W. (30 September 1987). "The history of Halley's Comet". Philosophical Transactions of the Royal Society of London. Series A, Mathematical and Physical Sciences. 323 (1572): 349–367. Bibcode:1987RSPTA.323..349H. doi:10.1098/rsta.1987.0091. ISSN 0080-4614. S2CID 123592786.
  103. ^ Chisholm, Hugh, ed. (1911). "Adams, John Couch" . Encyclopædia Britannica. Vol. 1 (11th ed.). Cambridge University Press. pp. 177–178.
  104. ^ a b c Einstein, Albert (1916). "The Foundation of the General Theory of Relativity" (PDF). Annalen der Physik. 49 (7): 769–822. Bibcode:1916AnP...354..769E. doi:10.1002/andp.19163540702. Retrieved 3 September 2006.
  105. ^ Lorentz, H. A. (1900). "Considerations on Gravitation" (PDF). Proceedings of the Royal Netherlands Academy of Arts and Sciences (KNAW). 2: 559–574.
  106. ^ Thompson, Silvanus P. (2019). "Lord Kelvin". International Electrotechnical Commission. Retrieved 16 October 2019.
  107. ^ Andrzej, Stasiak (2003). "Myths in science". EMBO Reports. 4 (3): 236. doi:10.1038/sj.embor.embor779. PMC 1315907.
  108. ^ Haber, Heinz (1967) [1965]. "Die Expansion der Erde" [The expansion of the Earth]. Unser blauer Planet [Our blue planet]. Rororo Sachbuch [Rororo nonfiction] (in German) (Rororo Taschenbuch Ausgabe [Rororo pocket edition] ed.). Reinbek: Rowohlt Verlag. p. 52. Bibcode:1967ubp..book.....H. Der englische Physiker und Nobelpreisträger Dirac hat ... vor über dreißig Jahren die Vermutung begründet, dass sich das universelle Maß der Schwerkraft im Laufe der Geschichte des Universums außerordentlich langsam, aber stetig verringert." English: "The English physicist and Nobel laureate Dirac has ..., more than thirty years ago, substantiated the assumption that the universal strength of gravity decreases very slowly, but steadily over the course of the history of the universe.
  109. ^ a b "Big Bang's afterglow shows universe is 80 million years older than scientists first thought". The Washington Post. Archived from the original on 22 March 2013. Retrieved 22 March 2013.
  110. ^ Bondi, H. (1957). "Negative mass in general relativity". Reviews of Modern Physics. 29 (3): 423–428. Bibcode:1957RvMP...29..423B. doi:10.1103/revmodphys.29.423.
  111. ^ Einstein, Albert (1912). "Lichtgeschwindigkeit und Statik des Gravitationsfeldes". Annalen der Physik (in German). 38 (7): 355–369. Bibcode:1912AnP...343..355E. doi:10.1002/andp.19123430704.
  112. ^ Einstein, Albert (1912). "Zur Theorie des statischen Gravitationsfeldes". Annalen der Physik (in German). 38 (7): 443. Bibcode:1912AnP...343..443E. doi:10.1002/andp.19123430709.
  113. ^ a b Einstein, A. and Grossmann, M. (1913), Zeitschrift für Mathematik und Physik 62, 225
  114. ^ Walter, S. (2007). Renn, J. (ed.). "Breaking in the 4-vectors: the four-dimensional movement in gravitation, 1905–1910" (PDF). The Genesis of General Relativity. 3. Berlin, Germany: 193–252. Bibcode:2007ggr..conf..193W.
  115. ^ Nordström, G (1912). "Relativitätsprinzip und Gravitation". Physikalische Zeitschrift (in German). 13: 1126.
  116. ^ a b Nordström, G. (1913). "Zur Theorie der Gravitation vom Standpunkt des Relativitätsprinzips". Annalen der Physik (in German). 42 (13): 533. Bibcode:1913AnP...347..533N. doi:10.1002/andp.19133471303.
  117. ^ Pais, Abraham (2005). Subtle is the Lord: The Science and Life of Albert Einstein. New York: Oxford University Press. ISBN 978-0-19-152402-8. OCLC 646798828.
  118. ^ Einstein, Albert; Fokker, A. D. (1914). "Die Nordströmsche Gravitationstheorie vom Standpunkt des absoluten Differentkalküls". Annalen der Physik (in German). 44 (10): 321–328. Bibcode:1914AnP...349..321E. doi:10.1002/andp.19143491009.
  119. ^ Abbott, Benjamin P.; et al. (LIGO Scientific Collaboration and Virgo Collaboration) (2016). "Observation of Gravitational Waves from a Binary Black Hole Merger". Phys. Rev. Lett. 116 (6): 061102. arXiv:1602.03837. Bibcode:2016PhRvL.116f1102A. doi:10.1103/PhysRevLett.116.061102. PMID 26918975. S2CID 124959784.
  120. ^ Abbott, B. P.; Abbott, R.; Abbott, T. D.; Acernese, F.; Ackley, K.; Adams, C.; Adams, T.; Addesso, P.; Adhikari, R. X.; Adya, V. B.; Affeldt, C.; Afrough, M.; Agarwal, B.; Agathos, M.; Agatsuma, K.; Aggarwal, N.; Aguiar, O. D.; Aiello, L.; Ain, A.; Ajith, P.; Allen, B.; Allen, G.; Allocca, A.; Altin, P. A.; Amato, A.; Ananyeva, A.; Anderson, S. B.; Anderson, W. G.; Angelova, S. V.; et al. (2017). "Multi-messenger Observations of a Binary Neutron Star Merger". The Astrophysical Journal Letters. 848 (2): L12. arXiv:1710.05833. Bibcode:2017ApJ...848L..12A. doi:10.3847/2041-8213/aa91c9. S2CID 217162243.
  121. ^ a b Randall, Lisa (2005). Warped Passages: Unraveling the Universe's Hidden Dimensions. Ecco. ISBN 9780060531089. ISBN.
  122. ^ Feynman, R. P.; Morinigo, F. B.; Wagner, W. G.; Hatfield, B. (1995). Feynman lectures on gravitation. Addison-Wesley. ISBN 978-0-201-62734-3.
  123. ^ Zee, A. (2003). Quantum Field Theory in a Nutshell. Princeton University Press. ISBN.

Sources