Weight
In physics, weight is the force exerted on an object by the gravitational attraction of a massive body, such as Earth, pulling it toward the body's center.[1] This force is distinct from mass, which measures the object's inertia or amount of matter and remains constant regardless of location; weight, however, varies depending on the strength of the local gravitational field.[2] The magnitude of weight W for an object of mass m is given by the formula W = mg, where g is the acceleration due to gravity, approximately 9.8 m/s² near Earth's surface.[1] Weight is a vector quantity, possessing both magnitude and direction (typically downward toward the gravitational source), and is measured in newtons (N) in the International System of Units (SI), reflecting its nature as a force.[2] In everyday language, "weight" is often misused to refer to mass, leading to confusion; for instance, a person's "weight" is commonly expressed in kilograms, but this actually denotes mass, with true weight being the corresponding gravitational force.[1] Weight decreases with increasing distance from the gravitational center—such as at higher altitudes or on other celestial bodies—and becomes zero in free fall or deep space, a condition known as weightlessness, where the object and its surroundings accelerate together under gravity.[2] Apparent weight, influenced by additional forces like those in elevators or aircraft, can differ from true gravitational weight, highlighting its dependence on the net force acting on the object.[1]Fundamental Concepts
Gravitational Definition
In physics, weight is defined as the gravitational force exerted on an object due to the attraction between its mass and the mass of a celestial body, such as Earth. This force arises from Newton's law of universal gravitation, which states that every particle in the universe attracts every other particle with a force proportional to the product of their masses and inversely proportional to the square of the distance between their centers. The magnitude of this force F_g is given by the equation F_g = G \frac{M m}{r^2}, where G is the universal gravitational constant ($6.67430 \times 10^{-11} \, \mathrm{m^3 \, kg^{-1} \, s^{-2}}), M is the mass of the attracting body (e.g., Earth), m is the mass of the object, and r is the distance between their centers.[3][4] Near the surface of Earth, this general expression simplifies to the weight W of an object, expressed as W = m g, where g is the local gravitational acceleration. The value of g is derived by substituting Earth's mass M (approximately $5.972 \times 10^{24} \, \mathrm{kg}) and radius r (approximately $6.371 \times 10^6 \, \mathrm{m}) into the universal gravitation formula, yielding g \approx 9.80665 \, \mathrm{m/s^2} at standard sea level. This approximation holds because r is nearly constant for objects on Earth's surface, making g effectively uniform for most practical purposes, though it varies slightly with latitude and altitude.[2][5] As a force, weight is a vector quantity, with magnitude m g and direction pointing toward the center of the attracting body, perpendicular to the local surface in the absence of other influences. This downward orientation explains why objects fall toward Earth and why weight opposes upward forces in equilibrium scenarios. The gravitational definition of weight emerged as the foundational concept in classical physics through Isaac Newton's work in the late 17th century, particularly in his Philosophiæ Naturalis Principia Mathematica (1687), where he unified terrestrial and celestial mechanics by positing gravity as a universal force. Prior to Newton, "weight" often referred interchangeably to mass or heaviness without a clear distinction from quantity of matter, but his formulation established weight explicitly as the measurable effect of gravitational attraction, influencing all subsequent mechanics until relativity.[6][7]Operational and ISO Definitions
In practical applications, the operational definition of weight is the magnitude of the force indicated by a weighing scale or balance when an object is at rest under the standard acceleration due to gravity, defined as g_n = 9.80665 \, \mathrm{m/s^2}. This reading represents the force required to support the object against gravity in a static equilibrium, ensuring reproducible measurements in engineering, manufacturing, and quality control contexts. Scales are calibrated to this standard value to account for variations in local gravity, providing a consistent operational measure expressed in newtons or, conventionally, converted to mass units via division by g_n.[5] The International Organization for Standardization (ISO) formalizes weight in ISO 80000-4:2019 as a specific type of force, namely the weight force \mathbf{F_g}, given by \mathbf{F_g} = m \mathbf{g}, where m is the mass of the body and \mathbf{g} is the local acceleration of free fall vector; its magnitude, denoted w = |\mathbf{F_g}|, has the SI unit of newton (N). For standardized purposes, ISO distinguishes the conventional weight as w_c = m g_n, using the fixed standard gravity g_n to enable uniform comparisons independent of location. The acceleration g is the effective local acceleration of free fall, which includes the gravitational attraction and the centrifugal effect due to Earth's rotation. The standard explicitly excludes the effect of atmospheric buoyancy from the weight definition. In legal metrology, this definition supports distinctions between true gravitational force and practical force measurements, where weight is treated as a vector quantity aligned with the local plumb line but quantified under controlled conditions.[8] In trade and commerce, "standard weight" refers to the assigned conventional mass value of calibration artifacts, such as weights used to verify scales, which are calibrated to balance against reference standards in air under defined environmental conditions. These values are determined assuming a reference air density of 1.2 kg/m³ and a brass density of 8000 kg/m³ for the reference weight, effectively incorporating buoyancy effects into the conventional mass without altering the underlying force definition. The International Organization of Legal Metrology (OIML) aligns with ISO through Recommendation D 28, defining conventional mass as the numerical value of the result of weighing in air, equivalent to the mass of the reference weight that balances the object, thus facilitating accurate trade transactions while adhering to ISO's force-based framework.[9] For non-gravitational influences, ISO 80000-4 addresses weight determination in contexts like buoyancy through notes on measurement practices, stipulating that apparent forces from air displacement must be corrected when deriving true mass from scale readings to isolate the gravitational weight. Buoyancy corrections in air are applied using formulas that account for the densities of the object and calibration standard, such as the conventional mass \mathrm{CM}_x = m_x \frac{1 - \rho_a}{1 - \rho_a / \rho_x} (approximating for standard air density \rho_a \approx 1.2 \, \mathrm{kg/m^3}), where \rho_x is the object's density. This is critical in high-precision metrology, where uncorrected buoyancy can introduce relative errors up to about 0.1% for objects with densities near that of air (e.g., \rho_x \approx 1 \, \mathrm{g/cm^3}).[10]Apparent Weight
Apparent weight refers to the normal force exerted by a supporting surface on an object or person, which can differ from the true gravitational weight due to the effects of acceleration in non-inertial reference frames. In such scenarios, the apparent weight is given by the equation W_{app} = m(g + a), where m is the mass, g is the acceleration due to gravity, and a is the acceleration of the frame relative to an inertial frame, taken positive in the upward direction.[11] This formulation arises from Newton's second law applied to the object in the accelerated frame, where the net force includes both gravity and the fictitious force due to acceleration.[12] A common example occurs in an elevator accelerating vertically. When the elevator accelerates upward with acceleration a > 0, the apparent weight increases, making the occupant feel heavier, as the normal force must provide the additional force to produce the net upward acceleration.[13] Conversely, during downward acceleration, such as when the elevator cable slows to stop, the apparent weight decreases. In free fall, where a = -g, the apparent weight becomes zero, resulting in weightlessness, as experienced by objects in orbit or during the drop phase of certain rides. Buoyancy in fluids further modifies apparent weight by introducing an upward buoyant force that opposes gravity. According to Archimedes' principle, the buoyant force F_b equals the weight of the fluid displaced by the object, so the apparent weight is W_{app} = mg - F_b = mg - \rho V g, where \rho is the fluid density and V is the submerged volume.[14] This reduction explains why submerged objects appear lighter when weighed on a scale in a fluid, a principle fundamental to hydrostatics and used in density measurements.[15] In aviation, apparent weight varies during maneuvers involving acceleration, such as turns or loops, where centripetal acceleration can multiply the effective gravitational force, leading to g-forces that increase the normal force on the pilot or passengers.[16] Similarly, in amusement rides like roller coasters, rapid changes in direction produce accelerations that alter apparent weight; at the bottom of a loop, upward acceleration increases it, while at the top, it may approach zero, simulating free fall.[17] These effects highlight how apparent weight depends on the dynamics of the supporting structure rather than gravity alone.Relation to Mass
Distinction Between Mass and Weight
Mass is a scalar quantity that represents the amount of matter in an object or the resistance of that object to changes in motion, known as inertia. In the International System of Units (SI), mass is measured in kilograms (kg), and it remains constant regardless of the object's location in the universe.[18][19] In contrast, weight is the gravitational force acting on an object's mass, which varies depending on the strength of the local gravitational field. For instance, an object with a given mass will have less weight on the Moon, where gravity is about one-sixth that of Earth's, compared to its weight on Earth.[20][19] This distinction arises because weight is a vector quantity, directed toward the center of the gravitational source, and its magnitude is determined by the product of the mass and the local acceleration due to gravity, expressed as W = m g, where m is mass and g is the gravitational acceleration.[21] While mass is an intrinsic property, weight is extrinsic and context-dependent, emphasizing that mass is the fundamental attribute from which weight derives.[22] Common misconceptions often blur this boundary, particularly in historical and educational contexts. Prior to Newtonian mechanics, "weight" was used interchangeably for both the quantity of matter and the downward force it experienced, leading to the erroneous view that mass and weight were identical.[23] In modern education, students frequently confuse the two, such as assuming weight is invariant or equating units like the pound-mass (a measure of mass) with the pound-force (a measure of force equivalent to the weight of one pound-mass under standard gravity).[24] These errors persist due to everyday language where "weight" colloquially means mass, but scientifically, they represent distinct physical concepts.[20]Units of Mass and Weight
In the International System of Units (SI), the base unit for mass is the kilogram (kg), defined by fixing the numerical value of the Planck constant h to exactly $6.62607015 \times 10^{-34} when expressed in the unit \mathrm{J \cdot s}, where the second and joule are defined previously.[25] The derived SI unit for weight, treated as a force, is the newton (N), defined as $1 \, \mathrm{N} = 1 \, \mathrm{kg \cdot m/s^2}, representing the force required to accelerate a 1 kg mass by 1 m/s².[18] Non-SI units remain in widespread use, particularly in the United States customary and British imperial systems. For mass, the pound-mass (lbm or lbm) is common, defined exactly as $0.45359237 \, \mathrm{kg}. The corresponding unit for weight is the pound-force (lbf), defined as the force exerted by standard gravity on a 1 lbm mass, equivalent to approximately $4.448222 \, \mathrm{N}.[26] Other mass units include the avoirdupois ounce (oz), equal to $1/16 lbm or exactly $28.349523125 \, \mathrm{g}, and the stone (st), used primarily in the UK for human body weight and defined as 14 lbm or approximately $6.35029318 \, \mathrm{kg}.[27] The following table provides key conversion factors between SI and selected non-SI units for mass and weight:| Mass Unit | Symbol | Equivalent in kg | Weight Unit | Symbol | Equivalent in N |
|---|---|---|---|---|---|
| Kilogram | kg | 1 | Newton | N | 1 |
| Pound-mass | lbm | 0.45359237 | Pound-force | lbf | 4.448222 |
| Ounce (avoirdupois) | oz | 0.028349523125 | - | - | - |
| Stone | st | 6.35029318 | - | - | - |
Historical Development
Newtonian Mechanics
Prior to Newton's work, the concept of weight was understood through Aristotelian physics, where heavy objects were thought to possess a natural tendency to move downward toward the center of the universe, their speed of fall proportional to their heaviness in a given medium.[29] This view treated weight as an intrinsic property driving natural motion, without reference to an underlying force or universal principle.[29] Isaac Newton's formulation of weight emerged from his synthesis of terrestrial and celestial mechanics, famously illustrated by the anecdote of an apple falling from a tree at Woolsthorpe Manor, which prompted him to consider why it fell toward Earth rather than ascending or deviating sideways.[30] Though this story, first recounted publicly by Voltaire over half a century after Newton's death, lacks direct confirmation from Newton himself, it symbolizes his insight into a unifying gravitational force.[30] In his seminal 1687 work, Philosophiæ Naturalis Principia Mathematica, Newton proposed the law of universal gravitation, stating that every particle attracts every other with a force proportional to the product of their masses and inversely proportional to the square of the distance between their centers: F \propto \frac{m_1 m_2}{r^2}.[31] Applied to an object of mass m near Earth's surface, this yields the weight as the gravitational force W = m g, where g is the acceleration due to gravity, approximately $9.8 \, \mathrm{m/s^2}, derived as g = \frac{G M}{R^2} with G the gravitational constant, M Earth's mass, and R its radius.[32] Newton integrated this with his second law of motion, F = m a, positing that the net force on an object equals its mass times acceleration.[33] For a freely falling body near Earth, the gravitational force provides the acceleration a = g, directly equating weight to m g and explaining why all objects fall at the same rate in vacuum, independent of mass.[33] He derived g's value by linking it to planetary motion, using Kepler's third law and centripetal acceleration for circular orbits to show that the same inverse-square force governs both apples and moons.[32] To measure g, Newton conducted pendulum experiments, observing that the period of a simple pendulum relates to length L and g via T = 2\pi \sqrt{\frac{L}{g}}, allowing computation of g from precise timing and length measurements; he noted variations with latitude due to Earth's rotation and oblateness.[34] These experiments, detailed in Principia Book 3, confirmed the uniformity of gravitational acceleration across locations.[31]Relativistic Physics
In relativistic physics, the concept of weight is fundamentally reshaped by Einstein's equivalence principle, which posits that the effects of gravity are locally indistinguishable from those of acceleration in a non-inertial frame. This principle establishes the equivalence between gravitational mass—the property that determines the strength of gravitational attraction—and inertial mass—the property that resists changes in motion—implying that both respond identically to gravitational fields. As a result, weight measurements, which in classical terms rely on the gravitational force acting on mass, become tied to the observer's local frame, where the sensation of weight arises from the acceleration required to counteract free fall.[35][36] In general relativity, weight is interpreted as the proper acceleration experienced by an object deviating from geodesic motion in curved spacetime, rather than a simple force between masses. A geodesic represents the straightest possible path in spacetime, akin to free fall under gravity; thus, an object at rest on Earth's surface feels weight due to the upward proper acceleration provided by the ground, which prevents it from following this geodesic. Gravitational time dilation further complicates this, as clocks—and by extension, measurements of acceleration and weight—run slower in stronger gravitational fields, affecting the perceived duration and magnitude of weight in precise applications. This observer-dependent nature means that weight is not an absolute property but varies with the local metric of spacetime.[36] Key historical milestones underscore these relativistic redefinitions of weight and gravity. Albert Einstein published the field equations of general relativity on November 25, 1915, in the Proceedings of the Prussian Academy of Sciences, providing the mathematical framework for understanding gravity as spacetime curvature that influences weight through acceleration. The theory's prediction of light deflection by gravity was confirmed during the solar eclipse of May 29, 1919, by expeditions led by Arthur Eddington, which measured starlight bending near the Sun by approximately 1.75 arcseconds—twice the Newtonian value—validating the equivalence principle's implications for gravitational effects. Modern applications, such as the Global Positioning System (GPS), require relativistic corrections for gravitational time dilation; satellite clocks, orbiting in weaker gravity, run faster by about 45 microseconds per day compared to Earth-based clocks, necessitating adjustments to maintain positional accuracy within meters, as unaccounted effects would accumulate errors of kilometers daily.[37][38] Unlike the Newtonian view of weight as an absolute force F = mg determined solely by mass and gravitational field strength, the relativistic perspective renders weight observer-dependent, arising from local proper acceleration in curved spacetime and influenced by time dilation. Newtonian mechanics serves as the low-speed, weak-field limiting case of general relativity, accurately describing weight under everyday conditions on Earth.Measurement Techniques
Principles of Weighing
The principle of weighing using a balance relies on achieving mechanical equilibrium through equal torques about the fulcrum. In a beam balance, the torque due to the weight of the unknown mass m_1 at distance L_1 from the fulcrum balances the torque from a known mass m_2 at distance L_2, expressed as m_1 g L_1 = m_2 g L_2, where g is the local gravitational acceleration.[39] This relationship shows that the balance determines mass ratios directly, with g canceling out, making the measurement independent of variations in gravitational field strength.[40] Spring scales operate on Hooke's law, which states that the restoring force F of a spring is proportional to its displacement x from equilibrium: F = -k x, where k is the spring constant. In weighing applications, the gravitational force mg stretches the spring by x = mg / k, and the scale is calibrated under standard g (typically 9.80665 m/s²) to read the weight directly as a force or equivalent mass.[41]/07%3A_Strength_and_Elasticity_of_the_Body/7.05%3A_Measuring_Weight) This calibration assumes linear elasticity within the spring's operating range, though deviations occur at high loads. Accuracy in weighing is influenced by environmental and operational factors, including air currents that induce buoyancy or aerodynamic torques, temperature variations causing thermal expansion or contraction in scale components, and magnetic fields that exert forces on ferromagnetic materials in the device.[42] These effects can introduce systematic errors, necessitating controlled conditions such as draft shields and temperature-stable environments for precise measurements.[43] Gravimetric weighing principles involve direct measurement of mass via gravitational force comparison or transduction, providing high precision for solids and calibrated volumes through weight differences.[44] Apparent weight adjustments may account for buoyancy in these methods./08%3A_Gravimetric_Methods)Types of Weighing Devices
Mechanical scales represent some of the earliest and simplest forms of weighing devices, relying on levers, beams, and counterweights to achieve balance without electrical components.[45] Beam balances, such as the two-pan equal-arm type, consist of a central fulcrum with identical arms suspending pans on each end, where the object is weighed against known standard masses to establish equilibrium.[46] These suspended-pan designs, also known as even-balance scales, were refined over centuries for applications in trade and science, offering reliability in environments without power sources.[46] Platform scales, a variant of mechanical beam systems, feature a flat platform supported by levers and springs that transmit the load to a dial indicator, enabling the weighing of larger items like crates or produce up to several hundred kilograms.[47] The historical evolution of mechanical weighing devices traces back to ancient civilizations, with the steelyard—a single-beam lever scale with a fixed fulcrum, sliding counterweight, and hook for suspending the object—originating in the ancient Near East around 3000 BCE and widely used by the Romans as a portable and efficient tool for commerce across the empire from the 1st century BCE.[48] This design, which allowed weighing up to 100 kilograms or more by adjusting the poise along a graduated beam, influenced subsequent European scales and persisted in various forms until the 19th century.[49] Electronic weighing devices have largely supplanted mechanical ones in modern applications due to their precision and speed, primarily utilizing load cells equipped with strain gauges to convert applied force into measurable electrical signals.[50] In these systems, strain gauges bonded to the load cell's structural element deform under weight, altering electrical resistance in a Wheatstone bridge circuit, which is then amplified and processed for output.[51] Digital readouts display the results instantly on LCD or LED screens, often with features like tare functions and data connectivity, supporting capacities from grams to tons in commercial settings.[50] For specialized needs, precision instruments extend the capabilities of both mechanical and electronic principles. Analytical balances, commonly used in laboratories, achieve resolutions as fine as 0.1 mg through enclosed chambers and electromagnetic force compensation in electronic models, ensuring minimal interference for weighing small samples like chemicals or pharmaceuticals up to 200 grams.[52] At the opposite end of the scale, industrial truck scales employ multiple high-capacity load cells—typically four to six, each rated 20 to 50 tons—mounted under a concrete or steel platform to weigh entire vehicles, facilitating logistics and compliance with transport regulations for loads exceeding 100 tons.[53] Advancements in precision have culminated in modern devices like quartz crystal microbalances, which use the piezoelectric properties of quartz resonators to detect mass changes at the nanogram level by monitoring shifts in resonant frequency, primarily for research in thin-film deposition and surface analysis.[54] These instruments, developed in the late 20th century, mark a shift from macroscopic mechanical systems to nanoscale electronic detection, calibrated against standard gravitational references for traceability.[54]Variations Across Environments
Weight on Earth
Weight on Earth varies primarily due to the planet's oblate spheroid shape, rotational effects, distance from the center of mass, and local geological density variations, leading to differences in the acceleration due to gravity, denoted as g. These factors cause g to range from approximately 9.780 m/s² at the equator to 9.832 m/s² at the poles, a variation of about 0.5% across the surface.[55][56] The primary latitudinal variation arises from Earth's oblateness—flattened at the poles and bulging at the equator due to centrifugal forces from rotation—and the direct centrifugal reduction in effective gravity. At the equator, the centrifugal acceleration reaches a maximum of about 0.033 m/s² outward, perpendicular to the axis of rotation, reducing the net g by roughly 0.3%.[57] Combined with the greater distance from Earth's center at the equator (equatorial radius ~21 km larger than polar), this results in the minimum g of 9.780 m/s², while the poles, lacking centrifugal effects and closer to the center, experience the maximum of 9.832 m/s².[55] Intermediate latitudes follow a sinusoidal pattern, with g decreasing toward the equator and increasing toward the poles.[56] As altitude increases above sea level, g decreases inversely with the square of the distance from Earth's center, following g \propto 1/r^2, where r is the radial distance. For heights up to 10 km, this yields an approximate linear decrease of -0.003 m/s² per kilometer, such that g at 10 km altitude is about 0.03 m/s² less than at sea level.[56] This effect is most relevant in aviation and mountaineering contexts, where the change remains small but measurable. Local gravitational anomalies, typically on the order of milligals (10^{-5} m/s²), arise from subsurface density variations due to geological features. For instance, regions with excess mass, such as dense ore deposits, produce positive anomalies, increasing g locally. Mountain ranges like the Himalayas exhibit positive free-air anomalies due to the mass of the topography itself, which provides a local enhancement relative to a hypothetical flat surface at the same elevation. However, Bouguer anomalies, which correct for the mass of the overlying rock, are large and negative (e.g., -180 to -450 mGal over the Himalayas) because the mountains are isostatically compensated by lower-density crustal roots beneath them. Consequently, the net effective g at mountain summits remains lower than at sea level due to the dominant effect of altitude. Sedimentary basins or low-density crustal features cause negative anomalies. These anomalies are mapped using gravimetry and inform geophysical surveys.[58][59]Weight on Other Celestial Bodies
The weight of an object on other celestial bodies is given by the Newtonian formula W = m \cdot g_{\text{body}}, where m is the invariant mass of the object and g_{\text{body}} is the local surface gravitational acceleration, which varies based on the body's mass and radius.[60] This contrasts with Earth's standard g \approx 9.81 \, \text{m/s}^2, providing a baseline for comparisons. On the Moon, surface gravity is approximately $1.62 \, \text{m/s}^2, resulting in an object's weight being about one-sixth of its Earth value.[61] Mars has a surface gravity of roughly $3.71 \, \text{m/s}^2, yielding a weight about 0.38 times that on Earth.[62] Jupiter's equatorial surface gravity is about $24.8 \, \text{m/s}^2, more than 2.5 times Earth's, though it decreases toward the poles due to the planet's rapid rotation.[62] Surface gravities for major solar system bodies are summarized below, with values relative to Earth for context:| Body | Surface Gravity (m/s²) | Relative to Earth |
|---|---|---|
| Moon | 1.62 | 0.165 |
| Mercury | 3.70 | 0.377 |
| Venus | 8.87 | 0.905 |
| Earth | 9.81 | 1.000 |
| Mars | 3.71 | 0.378 |
| Jupiter (equator) | 24.79 | 2.528 |
| Saturn (equator) | 10.44 | 1.065 |
| Uranus | 8.87 | 0.905 |
| Neptune | 11.15 | 1.137 |
| Pluto | 0.62 | 0.063 |