Fact-checked by Grok 2 weeks ago

Force

In physics, a force is a push or pull upon an object resulting from the object's interaction with another object, which can cause the object to accelerate, decelerate, or change direction. Forces always arise from interactions and cease when the interaction ends. A force is a quantity, characterized by both magnitude and direction, and is measured in the SI unit of the , defined as the force required to accelerate a mass of one kilogram at a rate of one meter per second squared. When multiple forces act on an object, their effects are determined by vector addition; balanced forces result in no net change in motion, while unbalanced forces produce acceleration proportional to the and inversely proportional to the object's mass, as described by Newton's second law (F = ma). The concept of force is foundational to , encapsulated in Isaac Newton's three laws of motion published in 1687. Newton's states that an object at rest remains at rest, and an object in uniform motion continues in a straight line unless acted upon by an unbalanced force, establishing the idea of . The third law asserts that for every action, there is an equal and opposite reaction, meaning forces always occur in pairs. Forces are classified into contact forces, which require physical touch between objects (such as friction, tension, normal force, and applied force), and non-contact forces (or field forces), which act at a distance without direct interaction (such as gravitational, electric, and magnetic forces). At the most fundamental level, all forces in nature are manifestations of four interactions: the gravitational force (weakest, responsible for planetary motion and weight), the electromagnetic force (governing electricity, magnetism, and chemical bonds), the weak nuclear force (involved in radioactive decay), and the strong nuclear force (binding atomic nuclei). These principles extend beyond classical physics into relativity and quantum mechanics, where force concepts are reformulated but remain central to understanding physical phenomena.

Historical Development

Pre-Newtonian Concepts

In , Aristotle conceptualized force primarily as an efficient cause or "mover" that initiates and sustains motion in bodies, distinguishing between natural and unnatural types of movement. Natural motion arises from inherent qualities within an object, such as heaviness compelling to move downward toward their natural place at the center of the , or lightness driving air and fire upward. This view positioned force not as an abstract quantity but as a teleological agent aligned with an object's essence and purpose, where motion ceases once the body reaches its proper position unless interrupted. Aristotle further differentiated natural forces from violent or unnatural ones, which require an external agent to compel motion contrary to an object's inherent tendencies, such as throwing a stone upward against its heaviness. Violent motion was seen as temporary and dependent on continuous application of the mover, with the speed proportional to the force exerted and inversely to the resistance encountered, though without quantitative measurement. This qualitative framework influenced physics for centuries, emphasizing motion's directional and qualitative properties over precise dynamics. During the medieval period, scholars like Jean Buridan and advanced these ideas through the impetus theory, positing that a projected body acquires a temporary, force-like quality or "impetus" from the initial mover, which sustains motion until dissipated by external resistance such as air or gravity. Buridan described impetus as an impressed motive power that decreases over time, explaining continued without ongoing external force, while Oresme extended this to falling bodies by suggesting impetus could accelerate descent if aligned with natural heaviness. These developments refined Aristotelian distinctions by introducing a quasi-conserved quality to violent motion, bridging natural tendencies and external influences in qualitative terms. Earlier, applied mechanical principles to forces through his analysis of levers, demonstrating how balanced weights on a beam pivoted around a could achieve or amplify effort without invoking Aristotelian movers directly. In works like On the Equilibrium of Planes, he illustrated practical force applications, such as using longer arms to counterbalance heavier loads, providing empirical insights into that prefigured later . These examples highlighted force as a relational property in rigid systems, influencing medieval engineers in constructing devices like catapults.

Newtonian Formulation

In the late 17th century, synthesized earlier ideas on motion into a rigorous mathematical framework for understanding force, culminating in his seminal work , first published in 1687. This text marked a departure from qualitative philosophies, introducing force as a quantifiable entity central to and resolving longstanding debates in astronomy, particularly those surrounding planetary motion. Building briefly on pre-Newtonian concepts such as the medieval , which qualitatively explained sustained motion through an internal "impetus" acquired from an initial push, Newton shifted toward precise, mathematical descriptions. Newton's formulation drew key influences from contemporaries and predecessors, notably and . Galileo's work on , which posited that bodies in motion tend to continue uniformly unless acted upon, provided the foundation for 's concept of unaltered motion in the absence of external influences. In contrast, rejected Descartes's vortex theory, which explained celestial motions through mechanical whirlpools of subtle matter, favoring instead without such intermediaries. These influences allowed to integrate terrestrial and under a unified system. Central to Newton's approach was his definition of force as vis impressa, or impressed force: "An impressed force is an action exerted upon a , in order to change its , either of rest, or of uniform motion in a right line. This consists in only, and remains no longer in the body when the action is over." This conception framed not as an enduring property but as a transient cause of motion change, enabling mathematical treatment. In the Principia, Newton applied this to resolve debates on planetary motion by deriving Johannes Kepler's empirical laws—such as elliptical orbits and the equal areas rule—from his theory of universal gravitation, demonstrating that an inverse-square attractive between bodies accounts for these phenomena as approximations in a solar system with minimal perturbations.

Newtonian Mechanics

First Law of Motion

The first law of motion, also known as the law of , states that every body perseveres in its state of rest or of uniform motion in a right line, unless it is compelled to change that state by forces impressed upon it. This establishes that objects maintain their —whether zero (at rest) or constant in magnitude and direction—without the application of an external , introducing the concept of inertia as the inherent resistance of matter to changes in motion. Historically, this overturned the Aristotelian view that objects naturally come to rest due to an intrinsic tendency toward immobility, positing instead that rest and uniform motion are equivalent states requiring no sustaining force. Aristotle's framework described motion as requiring continuous application of force, with natural motion for earthly bodies being toward rest at the center of the , a notion that dominated physics for nearly two millennia until challenged by Newtonian in the late . By contrast, Newton's formulation shifted the paradigm to one where change in motion demands an impressed force, laying the groundwork for . The law holds specifically in inertial reference frames, defined as those where an isolated object experiences no acceleration and thus maintains constant velocity, serving as the foundational condition for applying Newtonian mechanics. Newton invoked the concept of absolute space—immobile and independent of observable bodies—to justify this, as illustrated in his rotating bucket experiment: when a bucket filled with water is spun, the water's surface concaves due to centrifugal effects, demonstrating rotation relative to absolute space even if no external reference is visible, thereby evidencing true inertial motion. Conceptually, this implies that a net force of zero (\vec{F}_{\text{net}} = 0) results in zero acceleration (\vec{a} = 0), preserving uniform motion or rest.

Second Law of Motion

The second law of motion states that the acting on an object is equal to the product of its and its , expressed as the vector equation \vec{F}_{\text{net}} = m \vec{a}. This law quantifies how an unbalanced force causes a change in the object's , building on the concept of where no net force results in constant motion. In vector form, both and are vectors, meaning the direction of the matches the direction of the , while the of the is inversely proportional to the for a given . For instance, if multiple forces act on an object, their sum determines the and thus the resulting . The law derives from the rate of change of , where \vec{p} is defined as times (\vec{p} = m \vec{v} for constant mass systems), so \vec{F}_{\text{net}} = \frac{d\vec{p}}{dt}. When remains constant, this simplifies to \vec{F}_{\text{net}} = m \frac{d\vec{v}}{dt} = m \vec{a}. This formulation implies the SI unit of force, the (N), defined as the force required to accelerate a 1 mass by 1 m/s², or equivalently 1 N = 1 ·m/s². A practical example is a constant force pushing a 10 horizontally across a frictionless surface, resulting in an acceleration of 2 m/s² in the direction of the push, as \vec{F}_{\text{net}} = 10 \, \text{[kg](/page/KG)} \times 2 \, \text{m/s}^2 = 20 \, \text{N}. Here, represents the object's resistance to acceleration (), distinct from , which is the gravitational force mg where g \approx 9.81 \, \text{m/s}^2 on .

Third Law of Motion

Newton's third law of motion states that for every action, there is always an equal and opposite reaction, or more precisely, the mutual actions of two bodies upon each other are always equal in magnitude and directed to contrary parts. This principle, articulated by in his (1687), implies that if body A exerts a force \mathbf{F}_{AB} on body B, then body B exerts an equal and opposite force \mathbf{F}_{BA} = -\mathbf{F}_{AB} on body A. These action-reaction pairs always occur simultaneously and act along the line connecting the two bodies, emphasizing the symmetry inherent in physical interactions. In analyzing systems of particles, the third law distinguishes between internal and external forces: internal forces are those action-reaction pairs between objects within the , while external forces originate from outside the . For an —free from external forces—the internal forces sum to zero because each pair cancels in magnitude and direction, resulting in no on the as a whole. This has direct implications for conservation: since the net external force is zero (referencing the second law's relation to ), the total of the remains constant over time. A classic example is the recoil of a : when a is fired forward by expanding gases, the experiences an equal backward force, propelling it in the opposite direction with equal in magnitude but opposite to that of the . Similarly, relies on the third law, as the expels hot exhaust gases rearward at high , generating an equal forward on the to achieve liftoff and in the of . A common misconception is that action-reaction forces "cancel" each other out, leading to no motion; however, since these forces act on different bodies, they do not cancel for individual objects but instead cause each to accelerate according to its mass (as per the second ). For instance, in the gun recoil example, the forward force accelerates the lightweight rapidly, while the equal backward force accelerates the heavier more slowly, producing noticeable kickback without violating the .

Definition and Properties

In Newtonian mechanics, force is defined as a quantity that represents the capable of producing a change in the motion of a body, characterized by its , , and point of application. This definition stems from Isaac Newton's second law, where is proportional to the rate of change of , but fundamentally, it quantifies the push or pull exerted on an object. The nature of allows it to be represented in Cartesian coordinates as \vec{F} = (F_x, F_y, F_z), where each component corresponds to the along the respective . Forces are broadly classified into two types: contact forces, which arise from direct physical interaction between objects, such as or , and field forces (also known as action-at-a-distance forces), which act through empty space without requiring , exemplified by gravitational or electrostatic forces. forces fundamentally originate from electromagnetic interactions at the atomic level, while field forces are mediated by pervading . A key property of forces in Newtonian is their additivity, governed by the , which states that multiple forces acting on a combine vectorially to yield the , such that \vec{F}_{net} = \sum \vec{F}_i. This principle ensures that the effect of concurrent forces is equivalent to the single obtained by . Historically, the concept of force evolved from Newton's notion of "impressed force" in his (1687), described as an action exerted on a body to alter its state of rest or uniform rectilinear motion, measured by the change in momentum it produces. This marked a shift from pre-Newtonian ideas like impetus to a quantitative, mathematical framework. In modern extensions, particularly , force is generalized to the four-force tensor f^\mu = \frac{d p^\mu}{d \tau}, where p^\mu is the and \tau is ; in the low-velocity Newtonian limit, this reduces to the classical three-force \vec{F} = m \vec{a}.

Combining and Applying Forces

Force Equilibrium

Force equilibrium occurs when the vector of all forces on an object is zero, resulting in zero of the object's , as derived directly from Newton's second law of motion, \vec{F}_{net} = m \vec{a}, where if \sum \vec{F} = 0, then \vec{a} = 0. This condition holds for both objects at rest and those moving with constant velocity in the absence of net forces. To analyze force equilibrium, free-body diagrams (FBDs) are constructed, which isolate the object and depict all external forces on it as , including their magnitudes and directions, while excluding internal forces or forces exerted by the object itself. These diagrams facilitate the application of conditions by allowing the resolution of forces into components and verification that their is zero in each direction. Static equilibrium describes the case where an object remains at rest relative to an inertial frame, with \sum \vec{F} = 0 ensuring no translational motion and, briefly, no net for rotational stability (detailed in ). A example is a balanced , such as a beam balance where equal masses on either side produce equal and opposite gravitational forces and support reactions, keeping the beam horizontal without tipping. Dynamic equilibrium, in contrast, applies to objects moving at constant velocity, where \sum \vec{F} = 0 maintains uniform motion without acceleration, as seen in an idealized scenario of a sliding on frictionless ice, opposed only by balanced air resistance or negligible forces. can be translational, focusing on zero for , or rotational, requiring zero to prevent , though the latter is analyzed separately in torque discussions.

Net Force and Resultants

When multiple forces act on an object, their combined effect is described by the , or , which is the of all individual forces: \vec{R} = \sum_i \vec{F}_i. This determines the overall motion of the object, as it alone produces the same as the system of forces would. Forces, being vectors, add according to the , first articulated by in his . In this method, two forces \vec{F_1} and \vec{F_2} are represented as adjacent sides of a , with the \vec{R} as the diagonal from the common origin. For more than two forces, the polygon law extends this : forces are drawn head-to-tail in a closed polygon if in equilibrium, or with the as the from start to end point otherwise. To compute the resultant analytically, forces are resolved into components along perpendicular axes, such as Cartesian coordinates: R_x = \sum F_{ix}, R_y = \sum F_{iy}, and then |\vec{R}| = \sqrt{R_x^2 + R_y^2}, with direction given by \tan^{-1}(R_y / R_x). In applications like , the is the vertical gravitational force mg downward (assuming no air resistance), while the component is zero; this unbalanced resultant causes constant and downward of g = 9.8 \, \mathrm{m/s^2}. For an object on a frictionless at \theta to the , the to the plane is the component of mg \sin \theta down the incline, leading to a = g \sin \theta along the surface. Graphical methods, such as force polygons, are particularly useful in for visualizing in concurrent force systems; each force is scaled and drawn sequentially head-to-tail, with the as the straight-line closing the . Unlike force equilibrium, where the is zero and no occurs, a non-zero produces \vec{a} = \vec{R}/m in the direction of \vec{R}, as per Newton's second law.

Examples of Forces

Gravitational Force

The gravitational force is a fundamental attractive interaction between any two objects with , acting as a long-range, that requires no medium for propagation. This force, often simply called , is responsible for phenomena ranging from the falling of objects on to the motion of celestial bodies. It is always attractive and directed along the line connecting the centers of the two masses, with no repulsive counterpart in . Isaac Newton formulated the law of universal gravitation in his 1687 work , stating that the magnitude of the gravitational F_g between two point masses m_1 and m_2 separated by a distance r is given by F_g = G \frac{m_1 m_2}{r^2}, where G is the . The acts toward the center of each mass, and the law follows an inverse-square dependence on distance, meaning it weakens rapidly as separation increases. This formulation unified terrestrial and , explaining why the same causes apples to fall and planets to . The value of G was first experimentally determined by in 1798 through a torsion balance apparatus that measured the subtle between lead spheres, yielding G \approx 6.74 \times 10^{-11} \, \mathrm{m^3 \, kg^{-1} \, s^{-2}} (modern refinements confirm $6.67430 \times 10^{-11} with high precision). Cavendish's method isolated gravitational effects from other forces, providing the first quantitative link between and without relying on astronomical observations. Applications of gravitational force include planetary orbits, where it provides the necessary for stable elliptical paths as described by Newton's laws and Kepler's earlier empirical rules. Near Earth's surface, the force manifests as , defined as W = m g, where g is the local , approximately $9.81 \, \mathrm{m/s^2}. This weight arises from Earth's pulling objects downward. In representation, the force on a test m due to a larger of M is F_g = m g, with g = G \frac{M}{r^2} pointing toward the body's center, treating the field as a vector quantity independent of the test . The inverse-square law assumes point masses or spherically symmetric distributions, where the force outside a uniform sphere equals that of a point mass at its center; deviations occur for non-spherical or extended bodies, such as tidal effects from irregular shapes. This limitation highlights the law's idealization for practical calculations in astrophysics and engineering.

Electromagnetic Force

The electromagnetic force is a fundamental interaction that acts between electrically charged particles, manifesting as both attractive and repulsive effects depending on the charges involved. It combines electric forces, which arise between stationary charges, and magnetic forces, which affect moving charges or currents. Unlike the always-attractive gravitational force, the electromagnetic force can produce both attraction and repulsion, and it is vastly stronger at short ranges, dominating atomic and molecular structures. This force is responsible for phenomena ranging from chemical bonding to the operation of electrical devices. The electric component of the force is governed by , first experimentally established by in 1785 using a torsion balance to measure attractions and repulsions between charged objects. The law states that the magnitude of the electrostatic force F_e between two point charges q_1 and q_2 separated by a distance r is F_e = k \frac{|q_1 q_2|}{r^2}, where k = \frac{1}{4\pi\epsilon_0} \approx 8.99 \times 10^9 \, \mathrm{N \cdot m^2 / C^2} is Coulomb's constant, with \epsilon_0 being the . The force is directed along the line joining the charges and is repulsive for like charges (both positive or both negative) and attractive for unlike charges, as demonstrated in Coulomb's memoirs where he verified the inverse-square dependence through precise measurements. For example, the electrostatic repulsion between two s separated by atomic distances contributes to the stability of electron shells in atoms. The electric force is mediated by the \mathbf{E}, defined as the force per unit charge, which for a point charge is radial and falls off as $1/r^2. The magnetic component arises when charges are in motion, producing a force perpendicular to both the and the . This is captured in the magnetic force term \mathbf{F}_m = q (\mathbf{v} \times \mathbf{B}), where q is the charge, \mathbf{v} its , and \mathbf{B} the strength. The full electromagnetic force on a , known as the , combines both aspects: \mathbf{F} = q (\mathbf{E} + \mathbf{v} \times \mathbf{B}), formulated by in his 1895 theory of electromagnetic phenomena in moving bodies to explain effects on charged particles in fields. The magnetic force does no work on the particle since it is always perpendicular to the velocity, but it can alter the direction of motion, as in the deflection of charged particles in . The magnetic field \mathbf{B} mediates this interaction, with lines of force indicating its direction and strength. A practical example is in electric motors, where the on current-carrying conductors in a generates , causing rotation; for instance, in a simple , the interaction between the stator's field and the rotor's current produces continuous up to several newton-meters in typical designs. In the 1860s, James Clerk Maxwell unified these electric and magnetic forces into a coherent theory of by demonstrating that varying produce and vice versa, leading to the propagation of electromagnetic waves. His seminal paper introduced the concept of a pervasive governed by four equations (now known as ), which describe how \mathbf{E} and \mathbf{B} fields interact and mediate forces without direct contact between charges, analogous to the but with dynamic wave-like properties. This unification revealed that itself is an electromagnetic wave, with electric and magnetic fields oscillating perpendicular to the direction of propagation.

Contact Forces

Contact forces are those that arise from the physical between two objects in direct , as opposed to action-at-a-distance forces like or . In , these forces are essential for describing everyday phenomena such as pushing, pulling, or sliding objects. They typically act at the surface of and can be resolved into components and to that surface. The normal force is the component of the perpendicular to the surface of interaction, acting to prevent objects from penetrating each other. For an object on a surface, the normal force equals the object's , balancing the gravitational force. On an with angle θ, the normal force is given by N = mg \cos \theta, where m is the and g is the . Friction is the tangential component of the that opposes relative motion between surfaces. Static friction acts to prevent motion and has a maximum of f_s \leq \mu_s N, where \mu_s is the of static friction. Once motion begins, kinetic takes over, with magnitude f_k = \mu_k N, where \mu_k is the of kinetic friction, typically less than \mu_s. Tension is the pulling force transmitted along a flexible connector, such as a or , acting equally in all directions along its . For a mass m hanging in equilibrium from a massless , the equals , T = mg. Another example is the force, which arises from the elastic deformation of a and follows : F = -kx, where k is the spring constant and x is the displacement from . This restoring force is directed toward the equilibrium position. In , contact forces distributed over materials lead to , defined as per unit area, \sigma = F/A, where A is the cross-sectional area. This concept quantifies internal forces in solids under load.

Derived Concepts

Torque and Rotational Force

, often denoted as \vec{\tau}, is the rotational equivalent of in , representing the tendency of a force to cause about a pivot point or . It is defined as the of the position \vec{r} from the axis to the point of force application and the \vec{F}, given by \vec{\tau} = \vec{r} \times \vec{F}. The of is \tau = r F \sin \theta, where r is the distance from the axis to the of the (the lever arm), F is the , and \theta is the angle between \vec{r} and \vec{F}. This formulation highlights that depends not only on the but also on its perpendicular distance from the , emphasizing the rotational effect over pure translation. In rotational equilibrium, the net torque on an object about a given must be zero, analogous to the condition for translational where the is zero. This is expressed as \sum \vec{\tau} = 0, ensuring no occurs. For instance, on a balanced , the torque from one person's weight equals the counterclockwise from the other, maintaining regardless of their linear positions as long as the moments balance about the . The rotational analog of Newton's second law of motion relates net torque to angular acceleration \alpha through the moment of inertia I, which quantifies an object's resistance to rotational change based on its mass distribution. The equation is \vec{\tau}_{\text{net}} = I \vec{\alpha}, where I has units of kg·m². This law predicts that greater net torque produces larger angular acceleration for a given I, similar to how net force affects linear acceleration. A classic example is a lever, where applying a force at the end of a long arm (large r) generates substantial torque with minimal force, amplifying rotational effect around the pivot. Torque connects to linear force concepts through angular momentum \vec{L}, defined for a rigid body as \vec{L} = I \vec{\omega}, where \vec{\omega} is the angular velocity. The time derivative of angular momentum equals the net torque, \vec{\tau}_{\text{net}} = \frac{d\vec{L}}{dt}, linking rotational dynamics to changes in rotational motion. In a seesaw scenario, unequal torques alter the angular momentum, causing rotation until balance is restored or motion continues.

Potential Energy from Forces

In physics, a conservative force is defined as one for which the work done by the force on an object moving between two points depends only on the initial and final positions, independent of the path taken. This property allows the work to be associated with a energy function U, such that the force is the negative of this potential: \mathbf{F} = -\nabla U. Examples include gravitational and forces, where can be stored and recovered without loss. The gravitational potential energy between two point masses m_1 and m_2 separated by distance r is given by U_g = -\frac{G m_1 m_2}{r}, where G is the . Near Earth's surface, for an object of mass m at height h above a reference level, this approximates to U_g = m g h, where g is the . For elastic deformations, such as a obeying \mathbf{F} = -k \mathbf{x} (where k is the and \mathbf{x} is the from ), the associated is U_s = \frac{1}{2} k x^2. This quadratic form arises from integrating the force over , representing the energy stored in the deformed material. The work-energy theorem states that the net work done on an object equals the change in its : W = \Delta K_E. For conservative forces alone, the work done is W = -\Delta U, linking changes in directly to variations. In contrast, non-conservative forces like do not store energy reversibly; instead, they dissipate as or other forms, preventing full recovery. For instance, kinetic converts motion into , reducing the system's .

Conservation Laws

Conservation of arises directly from Newton's third of motion, which states that for every action there is an equal and opposite reaction. Consider an of particles interacting via internal forces. The total \mathbf{P} of the system is defined as the vector sum \mathbf{P} = \sum m_i \mathbf{v}_i, where m_i and \mathbf{v}_i are the and of the i-th particle. The time derivative of the total is \frac{d\mathbf{P}}{dt} = \sum \mathbf{F}_i, where \mathbf{F}_i is the on the i-th particle. By the third , internal forces between particles come in equal and opposite pairs, so their contributions to \sum \mathbf{F}_i cancel out. For an with no external forces, \frac{d\mathbf{P}}{dt} = 0, implying that \mathbf{P} is . This applies to closed systems, such as collisions between particles where external influences are negligible. In elastic or inelastic collisions, the total before and after the remains the same, enabling predictions of post-collision velocities from conditions. For example, in a one-dimensional collision between two masses, the equation m_1 v_{1i} + m_2 v_{2i} = m_1 v_{1f} + m_2 v_{2f} holds, where subscripts i and f denote initial and final states. Conservation of energy in mechanical systems follows from the properties of conservative forces, which are those where the work done depends only on initial and final positions, not the taken. For such forces, a U can be defined such that the force is \mathbf{F} = -\nabla U. The work-energy theorem states that the change in KE = \frac{1}{2} m v^2 equals the negative change in , so \Delta KE + \Delta U = 0. Thus, the total E = KE + U remains constant for systems governed solely by conservative forces, like or electrostatic interactions. A deeper connection between symmetries and conservation laws is provided by , which asserts that every of the laws of physics corresponds to a . Specifically, in space—meaning the laws are invariant under shifts in position—implies conservation of linear momentum. This theorem, derived from the invariance of the action integral in , unifies the origins of these principles. These laws hold rigorously only for isolated systems in non-relativistic . In non-isolated systems, external forces alter the total or . Relativistic effects, as in , modify the definitions—momentum becomes \mathbf{p} = \gamma [m](/page/M) \mathbf{v} where \gamma = (1 - v^2/c^2)^{-1/2}—but persists in inertial for the ./04%3A_Dynamics/4.03%3A_Relativistic_Momentum)

Units and Measurement

SI Units

The SI unit of force is the newton (N), defined as the force required to accelerate a mass of one kilogram at a rate of one meter per second squared, expressed as $1 \, \mathrm{N} = 1 \, \mathrm{kg \cdot m/s^2}. This definition stems from Newton's second law of motion, establishing force as the product of mass and acceleration. The newton was named in honor of Sir Isaac Newton for his foundational work on mechanics and adopted as the SI unit in 1948 by the General Conference on Weights and Measures (CGPM). Prior to this, force units varied by system, but the newton's adoption standardized measurements globally. In other systems, the (dyn) from the centimeter-gram-second (CGS) system equals $10^{-5} \, \mathrm{N}, while the pound-force (lbf) in the system is approximately $4.44822 \, \mathrm{N}. These conversions facilitate , with $1 \, \mathrm{dyn} = 1 \, \mathrm{g \cdot cm/s^2} and $1 \, \mathrm{lbf} = 1 \, \mathrm{lb \cdot ft/s^2} (where 1 lb ≈ 0.453592 ). Force is commonly measured using devices such as spring scales, which rely on to indicate or , and dynamometers, which quantify or linear force in applications. These instruments are calibrated against standards to ensure accuracy. Following the 2019 revision of the , the is now defined exactly in terms of the (h = 6.62607015 \times 10^{-34} \, \mathrm{J \cdot s}), enhancing the precision of force measurements by anchoring them to fundamental constants rather than artifacts. This change eliminates uncertainties from physical prototypes, improving traceability for the in .

Dimensional Analysis

Dimensional analysis examines the fundamental units of physical quantities to ensure consistency in equations and derive relationships between variables without solving full differential equations. The dimension of force, derived from Newton's second law F = ma, is expressed as [F] = MLT^{-2}, where M represents , L , and T time. This formulation highlights force as a product of inertial and , providing a basis for scaling physical laws across different systems. The formalizes by stating that any physical relationship involving n variables with k independent dimensions can be reduced to n - k dimensionless groups, or π terms. In , this theorem applies to drag force F_d, which depends on fluid density \rho, velocity v, characteristic length (e.g., radius r), and kinematic viscosity \nu. The theorem yields the C_d = \frac{F_d}{\rho v^2 r^2} as one π term and the Re = \frac{r v}{\nu} as another, leading to C_d = f(Re), where f is an unknown function determined experimentally. For high s, C_d approaches a constant, implying F_d \propto \rho v^2 A, with A as the cross-sectional area, which aids in predicting aerodynamic forces without full simulations. Scaling laws in emerge from by comparing similar systems, where lengths, times, and masses are scaled by factors \lambda_L, \lambda_T, and \lambda_M. For geometrically similar structures, forces scale with the square of the linear (F \propto L^2) to being of , while weights scale with volume (W \propto L^3). This results in a strength-to-weight decreasing as L^{-1} or M^{-1/3}, explaining why larger animals require proportionally thicker bones to support their . Such ratios guide designs, like model testing, by ensuring dynamic similarity through matched dimensionless numbers. A classic example is the simple pendulum, where the period T depends on bob m, string l, and g. Assuming T = C m^\alpha l^\beta g^\gamma with C dimensionless, equating dimensions yields \alpha = 0, \beta = 1/2, and \gamma = -1/2, so T = C \sqrt{l/g}. The exponent \alpha = 0 demonstrates the period's independence from , as it cancels out in the underlying . Despite its utility, dimensional analysis has limitations: it reveals only the form of relationships and dimensionless groups but cannot determine numerical constants or the exact functional form of f in expressions like \pi_1 = f(\pi_2, \dots). For instance, in the pendulum case, C = 2\pi requires solving the , not just dimensions. Additionally, it assumes all relevant variables are identified, and superfluous or incomplete sets can lead to overly complex or inaccurate results.

Modern Revisions

Relativistic Mechanics

In , the concept of force is revised to account for the effects of high velocities approaching the , c. The relativistic \mathbf{p} of a particle with rest m and velocity \mathbf{v} is given by \mathbf{p} = \gamma m \mathbf{v}, where the \gamma = \frac{1}{\sqrt{1 - \frac{v^2}{c^2}}} with v = |\mathbf{v}|. This expression arises from the invariance of the interval and ensures that momentum transforms correctly under Lorentz transformations, preserving conservation laws in relativistic collisions. The relativistic force is defined as the rate of change of this with respect to \tau, the time measured by a clock moving with the particle: \mathbf{F} = \frac{d\mathbf{p}}{d\tau}. Here, \mathbf{F} is the , a whose spatial components in the particle's instantaneous correspond to the three-force experienced by the particle. This differs from the Newtonian \mathbf{F} = m \mathbf{a}, as the effective increases with due to \gamma, leading to anisotropic : parallel to \mathbf{v}, the scales as \gamma^{-3} a, while perpendicular components scale as \gamma^{-2} a. In the low-velocity limit where v \ll c, \gamma \approx 1 + \frac{1}{2} \frac{v^2}{c^2}, so \mathbf{p} \approx m \mathbf{v} and \frac{d\tau}{dt} \approx 1, recovering the Newtonian form \mathbf{F} = m \mathbf{a}. This relativistic formulation is essential in applications like particle accelerators, where protons or electrons are accelerated to speeds exceeding 0.999c. For instance, in the , the from superconducting magnets provides the to maintain circular orbits, but the required strength increases with \gamma to counteract the growing relativistic , enabling energies up to several TeV without exceeding c. Such facilities demonstrate how fails at high speeds, as Newtonian predictions would underestimate the energy needed and predict unbounded velocities. In , the notion of force undergoes further revision: gravitational effects are not true forces but manifestations of curvature caused by mass-energy. Free particles follow —the shortest paths in curved —rather than deviating under a gravitational force, as described by the and the geodesic equation. This geometric interpretation eliminates the need for a gravitational , with apparent forces arising from non-geodesic motion in accelerated frames.

Quantum Mechanics

In , force is not a directly quantity in the same deterministic sense as in , where it is defined as the time rate of change of . Instead, the concept of force emerges through values of observables, as encapsulated by the . This theorem states that the value of the force acting on a particle, \langle \mathbf{F} \rangle = -\left\langle \nabla V(\mathbf{r}) \right\rangle, equals the times the second time of the value of , \langle \mathbf{F} \rangle = m \frac{d^2 \langle \mathbf{r} \rangle}{dt^2}, linking to classical on average. The theorem demonstrates how can approximate classical behavior for values, particularly when wave packets are well-localized compared to variations in the potential V(\mathbf{r}). The Heisenberg uncertainty principle imposes fundamental limits on measuring force precisely, as force relates to changes in momentum over position. The principle asserts that the product of uncertainties in position and momentum satisfies \Delta x \Delta p \geq \hbar/2, preventing simultaneous knowledge of a particle's exact location and velocity, which in turn obscures exact force determination since \mathbf{F} = dp/dt requires tracking momentum evolution. This intrinsic indeterminacy means that attempts to measure force at quantum scales inevitably disturb the system, leading to probabilistic outcomes rather than definite trajectories. For instance, in atomic physics, interatomic forces arise from gradients of quantum potentials, such as the Coulomb potential in molecules, where the force on a nucleus is the negative gradient of the total potential energy surface derived from the electronic wave function. A striking example of quantum forces deviating from classical expectations is quantum tunneling, where particles traverse potential barriers that would be impenetrable classically. In , for example, nuclear force confines an within a , but the particle's allows a nonzero probability of tunneling through, effectively experiencing a force that permits escape without surmounting the barrier height. This probabilistic penetration highlights how quantum forces, derived from potential gradients, enable phenomena like in stars, where tunneling overcomes electrostatic repulsion between protons. The time evolution of quantum states, governed by the i \hbar \frac{\partial \psi}{\partial t} = \hat{H} \psi, incorporates force indirectly through the \hat{H} = \frac{\hat{\mathbf{p}}^2}{2m} + V(\mathbf{r}), where the potential term relates to the force via \mathbf{F} = -\nabla V. Solutions to this equation yield wave functions \psi(\mathbf{r}, t) whose probability densities |\psi|^2 evolve under the influence of these potential-derived forces, often resulting in stationary states or superpositions that defy classical intuition. In the , recovers Newtonian force laws through Bohr's , which requires that quantum predictions match classical results for large quantum numbers or macroscopic scales, such as highly excited atomic states where de Broglie wavelengths become negligible. This principle ensures continuity between quantum and classical descriptions, with the providing the mathematical bridge for force dynamics.

Field Theories

In , forces are conceptualized as arising from spatial variations, or gradients, in underlying that permeate space. For example, the gravitational force on a test m is expressed as \mathbf{F} = -m \nabla \phi, where \phi is the scalar satisfying \nabla^2 \phi = 4\pi G \rho, with G the and \rho the . This formulation shifts the description from direct action-at-a-distance to a continuous mediating the interaction, as developed in Newtonian and extended to with the ./02%3A_Review_of_Newtonian_Mechanics/2.14%3A_Newtons_Law_of_Gravitation) Quantum field theory (QFT) reconciles with by treating particles as quantized excitations of pervasive fields, while forces emerge from the exchange of virtual particles—transient, off-shell field disturbances that do not obey the usual energy-momentum relation. In (QED), the electromagnetic force between charged particles is mediated by virtual photons, excitations of the photon field, enabling precise calculations of interactions like Coulomb scattering. This bridges classical field concepts with , where the fields are operator-valued and governed by commutation relations. The formalism underpins both classical and quantum field theories by providing a to derive . The action S = \int \mathcal{L} \, d^4x is extremized, where the Lagrangian density \mathcal{L} typically takes the form of terms minus potential-like interactions, analogous to the non-relativistic L = T - V but generalized to relativistic invariants such as \mathcal{L} = \frac{1}{2} \partial_\mu \phi \partial^\mu \phi - V(\phi) for a . The Euler-Lagrange equations \partial_\mu \left( \frac{\partial \mathcal{L}}{\partial (\partial_\mu \phi)} \right) - \frac{\partial \mathcal{L}}{\partial \phi} = 0 yield the field equations, facilitating quantization via path integrals. A key advancement in QFT is Yang-Mills theory, which extends gauge invariance to non-Abelian Lie groups, introducing self-interacting vector fields crucial for non-Abelian forces. Proposed for isotopic spin conservation, it describes the in (QCD) via SU(3) gauge fields, where gluons—massless, charged mediators—undergo nonlinear interactions due to the non-commuting group structure, leading to phenomena like . The exemplifies QFT's unifying power, framing the electromagnetic, weak, and strong interactions within a single renormalizable based on the group SU(3)_C \times SU(2)_L \times U(1)_Y, with fermions as chiral representations and for mass generation. This structure, developed through electroweak unification, accurately predicts scattering processes and particle masses, serving as the cornerstone of modern .

Fundamental Interactions

Strong and Weak Nuclear Forces

The strong nuclear force governs the interactions among quarks and is essential for the structure of hadrons and atomic nuclei. At the level, it binds quarks within protons, neutrons, and other hadrons through the exchange of gluons, which are massless vector bosons carrying , as described by (QCD). QCD, a non-Abelian based on the SU(3) color group, was theoretically established in the early 1970s, with the key discovery of —where the strong coupling constant decreases at high energies—demonstrated by , , and Hugh Politzer. This property allows perturbative calculations at short distances while explaining quark confinement at larger scales. The strong force operates over an extremely short range of approximately $10^{-15} m, the scale of nuclear dimensions, beyond which it drops off rapidly due to color confinement. Within this range, it is roughly 100 times stronger than the electromagnetic force, enabling it to dominate and overcome the Coulomb repulsion between positively charged protons in the nucleus. The residual strong force between composite nucleons (protons and neutrons) arises from the exchange of virtual pions and is modeled by the Yukawa potential, which takes the approximate form V(r) \propto \frac{e^{-r}}{r} in natural units. This potential captures the attractive, short-ranged nature of the nuclear binding. Additionally, the strong interaction exhibits approximate isospin symmetry, an SU(2) invariance that treats up and down quarks (and thus protons and neutrons) nearly equivalently, owing to their small mass difference of about 3 MeV. This symmetry underlies the near-degeneracy in masses of mirror nuclei and simplifies models of nuclear structure./08%3A_Symmetries_of_the_theory_of_strong_interactions/8.01%3A_The_First_Symmetry_-_Isospin) In atomic nuclei, the strong force provides the primary binding mechanism, with average binding energies per nucleon reaching up to 8-9 MeV near iron-56, as revealed by the nuclear binding energy curve. This curve, derived from mass defect measurements, shows that lighter nuclei gain stability through fusion (releasing energy by moving toward higher binding per nucleon), while heavier nuclei release energy via fission (splitting into fragments with higher average binding). For instance, the fusion of hydrogen into helium in stars or the fission of uranium-235 in reactors both exploit these strong-force-mediated bindings to liberate vast amounts of energy, far exceeding chemical reactions./University_Physics_III_-Optics_and_Modern_Physics(OpenStax)/10%3A__Nuclear_Physics/10.03%3A_Nuclear_Binding_Energy) The weak nuclear force, in contrast, mediates flavor-changing processes at an even shorter range of about $10^{-18} m and is responsible for radioactive and other transformations that alter particle identities. It is carried by the massive ^\pm and ^0 bosons, with masses around 80-91 GeV/c^2, which were experimentally confirmed at in 1983. In beta-minus decay, for example, a in a emits a ^- boson, transforming into an (yielding a proton) while the ^- decays into an and antineutrino; this charged-current interaction changes the particle's flavor and charge. The neutral-current process, mediated by the boson, allows flavor-preserving scattering but contributes to processes like interactions. A hallmark of the weak force is its violation of parity conservation, meaning it distinguishes between left- and right-handed particles, unlike and electromagnetic forces. This was experimentally verified in the 1957 , where beta electrons from the decay of polarized ^{60}Co nuclei were observed to be emitted preferentially opposite to the nuclear spin direction, indicating a preferred in the weak current. The result supported the vector-axial vector (V-A) theory of weak interactions proposed by Feynman and Gell-Mann, resolving puzzles in spectra and paving the way for the electroweak unification in the . The weak force's role in flavor changes also drives , such as the conversion of protons to neutrons in the rapid proton capture process.

Unification Attempts

One of the earliest successful attempts at unifying fundamental forces was the electroweak theory, which merges the electromagnetic and weak nuclear forces into a single framework. Proposed independently by in 1967 and in 1968, building on Sheldon Glashow's 1961 model, this theory posits that at high energies, the SU(2) × U(1) gauge symmetry governs both interactions, with the breaking the symmetry to yield the observed forces. Experimental confirmation came in 1973 through the discovery of weak neutral currents by the collaboration at , which observed neutrino interactions without charged lepton production, aligning with the theory's predictions. This unification, formalized in the Glashow-Weinberg-Salam model, forms a cornerstone of the and earned its progenitors the 1979 . Building on electroweak unification, Grand Unified Theories (GUTs) seek to incorporate the , combining the electromagnetic, weak, and strong interactions under a single gauge group at energies around 10^16 GeV. The minimal GUT, SU(5), proposed by and in 1974, embeds the Standard Model's SU(3) × SU(2) × U(1) into SU(5), predicting that quarks and leptons unify in multiplets and leading to observable consequences like with a lifetime of about 10^31 to 10^34 years. An extension, the SO(10) model introduced by Georgi in 1975, uses the larger SO(10) group to naturally include right-handed neutrinos, enabling seesaw mechanisms for neutrino masses and further predicting modes. While these theories elegantly unify three forces and explain charge quantization, searches at detectors like have set lower limits on lifetimes exceeding 10^34 years, constraining but not ruling out minimal GUTs. Efforts to unify all four fundamental forces, including , have led to frameworks like and its extension, . In , developed in the 1970s and refined through the 1980s, fundamental particles and forces arise from vibrational modes of one-dimensional strings in 10 dimensions, with different vibrations corresponding to gravitons, photons, gluons, and other mediators. , conjectured by in 1995, unifies the five consistent superstring theories into an 11-dimensional framework incorporating membranes (branes), where forces manifest as low-energy approximations of higher-dimensional dynamics.00140-3.full) These approaches promise a quantum theory of but remain untested, as their predictions emerge at the Planck scale of 10^19 GeV. Significant challenges persist in these unification attempts, particularly the inclusion of , which resists quantization within standard theory approaches, leading to non-renormalizable infinities in perturbative . The exacerbates this, questioning why the weak force scale (around 246 GeV) is so much smaller than the Planck scale without , a issue that GUTs and address through mechanisms like or but without definitive resolution. Moreover, while the has triumphed in describing electromagnetic, weak, and strong interactions with extraordinary precision—predicting phenomena like the discovered in 2012—it falls short on , masses, and the nature of , which constitutes about 27% of the universe's mass-energy yet evades particles. Ongoing experiments at the LHC and neutrino observatories continue to probe these gaps, but a complete unification remains elusive.

References

  1. [1]
    The Meaning of Force - The Physics Classroom
    A force is a push or pull upon an object resulting from the object's interaction with another object. Whenever there is an interaction between two objects, ...
  2. [2]
    What is Force? | NIST - National Institute of Standards and Technology
    Sep 11, 2009 · When we push or pull on a body, we are said to exert a force on it. Forces can also be exerted by inanimate objects.
  3. [3]
    Force - HyperPhysics Concepts
    Force. One of the foundation concepts of physics, a force may be thought of as any influence which tends to change the motion of an object.
  4. [4]
    Aristotle's Natural Philosophy
    May 26, 2006 · A mover can effect a motion which is contrary to its own nature. Aristotle's example of such an unnatural mover is the lever, an object heavy ...Missing: heaviness lightness
  5. [5]
    On the Heavens by Aristotle - The Internet Classics Archive
    And by absolutely light I mean one which of its own nature always moves upward, by absolutely heavy one which of its own nature always moves downward, if no ...
  6. [6]
    cambCompChap5Science
    Start with the observation that earth and other heavy things go down. Thus there is a "natural motion" for heavy objects. But what of fire? It goes up. The ...
  7. [7]
    Aristotle - Galileo and Einstein
    For violent motion, Aristotle stated that the speed of the moving object was in direct proportion to the applied force. This means first that if you stop ...
  8. [8]
    Aristotle's Laws of Motion
    There is natural, violent, and local motion; rectilinear and circular motion; Speed is proportional to motive force, and inversely proportional to resistance.
  9. [9]
    [PDF] Aristotle's Physics: a Physicist's Look - PhilSci-Archive
    While violent motion is the accelerated motion of an object subject to a force. ... Aristotelian distinction between natural and violent motion.8. This view of ...
  10. [10]
    John Buridan - Stanford Encyclopedia of Philosophy
    May 13, 2002 · The theory of impetus probably did not originate with Buridan, but his account appears to be unique in that he entertains the ...
  11. [11]
    Medieval Theories of Causation - Stanford Encyclopedia of Philosophy
    Aug 10, 2001 · However, what is more interesting for us are the alternative causal accounts that Buridan and Oresme adopted: they both said that projectiles ...
  12. [12]
    [PDF] John Buridan's 14th century concept of momentum - arXiv
    In the 14th century the French thinker John Buridan developed a theory of motion that bears a strong resemblance to Newtonian momentum.
  13. [13]
    Archimedes' Law of the Lever
    This is the statement of the Law of the Lever that Archimedes gives in Propositions 6 and 7 of Book I of his work entitled On the Equilibrium of Planes.Missing: applications | Show results with:applications
  14. [14]
    [PDF] Archimedes, the Center of Gravity, and the First of Mechanics:
    Archimedes obtained the law of the lever in Proposition 6 of the first part ... Complement to the law of the lever: The downward force exerted by the ...
  15. [15]
    [PDF] The Mechanical Problems in the Corpus of Aristotle
    Our author is pre-Archimedes; he is happy merely to (correctly) state the principle of the lever; Archimedes rigorously proves it. Simi- larly our author ...
  16. [16]
    Newton's Philosophiae Naturalis Principia Mathematica
    Dec 20, 2007 · The definitions inform the reader of how key technical terms, all of them designating quantities, are going to be used throughout the Principia.
  17. [17]
    Kepler and the origins of pre-Newtonian mass - AIP Publishing
    Feb 1, 2017 · Impetus was the progenitor of momentum, which is a measure of motion, not a cause. Impetus Theory set the stage for the advances made by the ...
  18. [18]
    Newton's Philosophy
    Oct 13, 2006 · Newton's first law reflects Descartes's laws: it is a new version of the principle of inertia, one incorporating the concept of an impressed ...
  19. [19]
    Newton's Mathematical Principles of Natural Philosophy
    The vis insita, or innate force of matter, is a power of resisting, by which every body, as much as in it lies, continues in its present state, whether it be of ...
  20. [20]
    Newton's First Law of Motion - Richard Fitzpatrick
    Indeed, we can think of Newton's first law as the definition of an inertial frame: i.e., an inertial frame of reference is one in which a point object subject ...
  21. [21]
    [PDF] Classical Mechanics 1 (PDF)
    Newton's first law. Page 12. Let us start with the first law Newton wrote down ... This very naturally leads to Aristotle's laws of motion. It is in the ...
  22. [22]
    Frames of Reference and Newton's Laws - Galileo
    An inertial frame is defined as one in which Newton's law of inertia holds—that is, any body which isn't being acted on by an outside force stays at rest if it ...
  23. [23]
    Newton's bucket - MacTutor History of Mathematics
    Newton deduced from this thought experiment that there had to be something to measure rotation with respect to, and that something had to be space itself. It ...
  24. [24]
    5.3 Newton's Second Law – University Physics Volume 1
    Vector division is not defined, so m = F → net / a → cannot be performed. However, mass m is a scalar, so we can use the scalar form of Newton's second law, ...
  25. [25]
    Linear Momentum and Force – ISP209: The Mystery of the Physical ...
    Δ p = Δ ( m v ) . If the mass of the system is constant, then. Δ ( m v ) = m Δ v . So that for constant mass, Newton's second law of motion becomes. F net = Δ p ...
  26. [26]
    Newton's Laws - Student Academic Success
    On a flat surface, Newton's Second Law shows that objects only move with an unbalanced force. When pushing a box with 100 N ...
  27. [27]
    [PDF] Newton's Principia : the mathematical principles of natural philosophy
    ... force of gravity, the primary planets must also be carried round the sun by ... NATURALIS PRINCIPIA. MATHEMATICA, dedicated to the Royal Society, and ...
  28. [28]
    Newton's Laws - HyperPhysics Concepts
    Newton's third law: All forces in the universe occur in equal but oppositely directed pairs. There are no isolated forces; for every external force that acts on ...
  29. [29]
    4.4 Newton's Third Law of Motion: Symmetry in Forces
    6: Newton's third law of motion tells us that forces always occur in pairs of equal and opposite magnitude. Explain how the choice of the “system of interest” ...
  30. [30]
    Newton's Third Law of Motion: Symmetry in Forces
    Newton's third law of motion tells us that forces always occur in pairs of equal and opposite magnitude. Explain how the choice of the “system of interest” ...
  31. [31]
    [PDF] 6-3 Implication of Newton's Third Law: Momentum is Conserved
    Key idea for momentum conservation: Even if the momenta of individual parts of a system are not conserved, the momentum of the entire system is conserved ( ...
  32. [32]
    Conservation of momentum
    Newton's third law implies that the total momentum of a system of interacting objects not acted on by outside forces is conserved. The total momentum in the ...
  33. [33]
    According to the third Newton's law - UCSB Science Line
    Jun 23, 2017 · This is Newton's third law in action. The change in momentum of the gun is called recoil. Because the gun is much more massive compared to the ...
  34. [34]
    Propulsion Activity - NASA Glenn Research Center
    Propulsion systems are governed by Newton's Third Law: "For every action, there is an equal and opposite reaction." Rocket Propulsion: Thrust equals Mass Flow ...
  35. [35]
    [PDF] A Taxonomy of Misconceptions - ASU Modeling Instruction
    ” Hence, students find Newton's Third Law unreasonable, and they prefer some version of the dominance principle: In a conflict, the. “more forceful” exerts ...
  36. [36]
    Newton's Third Law
    When a gun or cannon is fired, hot exploding gases push the bullet or shell forward. The gun itself is pushed backwards. This is called "recoil". Here's a ...
  37. [37]
    5.1 Forces – University Physics Volume 1 - UCF Pressbooks
    Understand the definition of force; Identify simple free-body diagrams; Define the SI unit of force, the newton; Describe force as a vector. The study of motion ...
  38. [38]
    [PDF] 8.01 Classical Mechanics Chapter 7.1-7.3 - MIT OpenCourseWare
    Force is a vector quantity. The direction of the force on the standard body is defined to be the direction of the acceleration of the body. Thus... F ...
  39. [39]
    Newton's second law of motion
    An important corollary of Newton's second law is that force is a vector quantity. ... forces; it lies at the heart of many calculations in Newtonian mechanics.
  40. [40]
    [PDF] Physics 111: Mechanics Lecture 3 - NJIT
    ❑ May be a contact force or a field force. ▫ Contact forces result from physical contact between two objects. ▫ Field forces act between disconnected ...
  41. [41]
    [PDF] NewtoN's Laws of MotioN - Pearson
    4.1 shows, force is a vector quantity; you can push or pull a body in different directions. When a force involves direct contact between two bodies, such as a ...
  42. [42]
    12 Characteristics of Force - Feynman Lectures - Caltech
    For example, the law of response of a mass to a gravitational field is that the force is equal to the mass times the gravitational field; or, if there is also a ...
  43. [43]
    [PDF] Acceleration and Force in Special Relativity Howard E. Haber
    Mar 1, 2024 · In these notes, the properties of the velocity, momentum, force and acceler- ation four-vectors of special relativity are examined.
  44. [44]
    Equilibrium and Free-body Diagrams - Copyright
    A free-body diagram is a sketch showing the loads acting on a rigid body. These loads may be forces and moments. The loads may be internally or externally ...
  45. [45]
    12.2 Examples of Static Equilibrium – University Physics Volume 1
    The system is in static equilibrium when the beam does not rotate. It is balanced when the beam remains level.
  46. [46]
    5.2 Newton's First Law – University Physics Volume 1
    Equilibrium is achieved when the forces on a system are balanced. A net force of zero means that an object is either at rest or moving with constant velocity; ...
  47. [47]
    12.1 Conditions for Static Equilibrium – University Physics Volume 1
    We say that a rigid body is in equilibrium when both its linear and angular acceleration are zero relative to an inertial frame of reference.
  48. [48]
    5.1 Forces - University Physics Volume 1 | OpenStax
    ### Summary of Content from https://openstax.org/books/university-physics-volume-1/pages/5-1-forces
  49. [49]
  50. [50]
    Why do forces add vectorially? A forgotten controversy in the ...
    Apr 1, 2011 · The parallelogram of forces seems to have been widely recognized by Newton's day because both Pierre Varignon and Bernard Lamy stated it in the ...INTRODUCTION · II. THE DYNAMICAL... · IV. POISSON'S STATIC...Missing: source | Show results with:source
  51. [51]
    4.3 Projectile Motion - University Physics Volume 1 - OpenStax
    Sep 19, 2016 · Some examples include meteors as they enter Earth's atmosphere, fireworks, and the motion of any ball in sports. Such objects are called ...Missing: net | Show results with:net
  52. [52]
    OpenStax | Free Textbooks Online with No Catch
    **Summary of Net Force on Inclined Plane (No Friction) from OpenStax University Physics Volume 1, Section 5.4:**
  53. [53]
    [PDF] Charles-Augustin Coulomb First Memoir on Electricity and Magnetism
    In a memoir presented to the Academy, in 1784, I have determined from experiments the laws governing the torsional resistance of a filament of metal and I.
  54. [54]
    Electric field - HyperPhysics
    The electric field of a point charge can be obtained from Coulomb's law: The electric field is radially outward from the point charge in all directions. The ...
  55. [55]
    Magnetic forces - HyperPhysics
    Lorentz Force Law. Both the electric field and magnetic field can be defined from the Lorentz force law: The electric force is straightforward, being in ...
  56. [56]
    Brushed DC Motor Theory - Northwestern Mechatronics Wiki
    Feb 16, 2011 · Motor cutaway.png. Motor Physics. The forces inside a motor that cause the rotor to rotate are called Lorentz Forces. If an electron is moving ...
  57. [57]
    [PDF] A Dynamical Theory of the Electromagnetic Field
    A Dynamical Theory of the Electromagnetic Field. J. Clerk Maxwell. , 459-512, published 1 January 1865. 155. 1865. Phil. Trans. R. Soc. Lond. Email alerting ...
  58. [58]
    Maxwell's Equations - HyperPhysics Concepts
    Maxwell's equations represent one of the most elegant and concise ways to state the fundamentals of electricity and magnetism.
  59. [59]
    [PDF] Chapter 5 - Force and Motion - UF Physics Department
    which is at the heart of classical mechanics. We must note that Newton's laws ... Contact Forces: As the name implies these forces act between two objects.
  60. [60]
    Normal Force – Physics 131: What Is Physics? - Open Books
    The normal force is the force that keeps objects from passing through each other, like a book on a table, and is a contact force.
  61. [61]
  62. [62]
    Friction - HyperPhysics Concepts
    Static friction resistance will match the applied force up until the threshold of motion. Then the kinetic frictional resistance stays about constant. This plot ...
  63. [63]
    4.5 Normal, Tension, and Other Examples of Forces – College Physics
    Normal Force. Weight (also called force of gravity) is a pervasive force that acts at all times and must be counteracted to keep an object from ...
  64. [64]
    Mechanics of Materials: Stress - Boston University
    Stress is the measure of an external force acting over the cross sectional area of an object. Stress has units of force per area: N/m 2 (SI) or lb/in 2 (US).
  65. [65]
    Torque - Front Matter
    The torque exerted by a force F → acting on a rigid body about a given axis is , τ → = r → × F → , where r → is the perpendicular vector from the axis to the ...
  66. [66]
    [PDF] Rotational Motion 3 - Duke Physics
    Here is the general definition of the torque of a force about a given reference point: The vector r specifies the location, relative to the reference point, of ...
  67. [67]
    [PDF] Chapter 10 Dynamics of Rotational Motion 1 Torque
    The vector definition of the torque can be written as: ~τ = r × ~F. (the torque vector). How is the angle φ defined? It is the angle between r and ~F. Recall ...
  68. [68]
    TUTORIAL : TORQUE AND ROTATIONAL EQUILIBRIUM
    The sum of the torques must equal zero. All the torques must be calculated with reference to the same point designated as the fulcrum, and any point may be ...
  69. [69]
    Rotational dynamics
    torque = lever arm × force, ... While the seesaw is moving there is practically no torque on the seesaw, and it rotates with uniform angular velocity.
  70. [70]
    10.7 Newton's Second Law for Rotation – University Physics Volume 1
    If more than one torque acts on a rigid body about a fixed axis, then the sum of the torques equals the moment of inertia times the angular acceleration:.
  71. [71]
    Newton's Second Law for Rotation - Physics
    Newton's second law states that the angular acceleration is proportional to the net torque and inversely proportional to the moment of inertia. Newton's ...
  72. [72]
    Rotational Motion - Galileo
    A familiar example of a body having forces applied at different points is the seesaw. Let's say we have a seesaw carrying two children, one weighing 20 kg ...
  73. [73]
    Angular momentum
    The angular momentum L about one of the object's symmetry axis is the product of the object's moment of inertia I times its angular velocity ω about this ...
  74. [74]
    The Law of Conservation of Angular Momentum - Physics
    L = Iω · Angular momentum is a vector, pointing in the direction of the angular velocity. · If there is no net torque acting on a system, the system's angular ...
  75. [75]
    45. 7.4 Conservative Forces and Potential Energy - UH Pressbooks
    A conservative force is one for which work done by or against it depends only on the starting and ending points of a motion and not on the path taken. We can ...
  76. [76]
    [PDF] Lecture D8 - Conservative Forces and Potential Energy
    In multiple dimensions, the condition for a force field to be conservative is that it can be expressed as the gradient of a potential function. That is,. F = − ...
  77. [77]
    Gravitational Potential Energy - HyperPhysics
    where G is the gravitation constant, M is the mass of the attracting body, and r is the distance between their centers. This is the form for the gravitational ...
  78. [78]
    7.3 Gravitational Potential Energy – College Physics
    The equation Δ PE g = m g h applies for any path that has a change in height of h , not just when the mass is lifted straight up. (See Figure 2.) It is much ...
  79. [79]
    Elastic Potential Energy - HyperPhysics
    According to Hooke's law, the force required to stretch the spring will be directly proportional to the amount of stretch.
  80. [80]
    7.4 Conservative Forces and Potential Energy - UCF Pressbooks
    Use the work-energy theorem to show how having only conservative forces implies conservation of mechanical energy. ... The work-energy theorem states that the net ...
  81. [81]
    [PDF] 8. Conservative Forces and Potential Energy
    Therefore, we can define, for conservative forces, an associated potential energy that, for a given object, depends only on its location. In particular ...
  82. [82]
    7.5 Nonconservative Forces – College Physics - UCF Pressbooks
    Friction, for example, creates thermal energy that dissipates, removing energy from the system.
  83. [83]
    Conservative and non-conservative force-fields - Richard Fitzpatrick
    Hence, friction is an example of a non-conservative force, because it dissipates energy rather than storing it.
  84. [84]
    9.3 Conservation of Linear Momentum - University Physics Volume 1
    Sep 19, 2016 · Recall Newton's third law: When two objects of masses m 1 m 1 and m 2 m 2 interact (meaning that they apply forces on each other), the force ...
  85. [85]
    10 Conservation of Momentum - Feynman Lectures - Caltech
    So, here is another law, which might be stated: If two bodies have equal masses, as measured by equal velocities at one velocity, they will have equal masses ...
  86. [86]
    [PDF] Invariant Variation Problems
    Emmy Noether. M. A. Tavel's English translation of “Invariante Variationsprobleme,” Nachr. d. König. Gesellsch. d. Wiss. zu Göttingen, Math-phys. Klasse, 235 ...
  87. [87]
    1.4 Dimensional Analysis – University Physics Volume 1
    The dimension of any physical quantity expresses its dependence on the base quantities as a product of symbols (or powers of symbols) representing the base ...
  88. [88]
    [PDF] 5.5 Buckingham Pi theorem - MIT
    May 13, 2010 · First, toward the low-Reynolds-number end, the drag coefficient increases. Second, for high Reynolds numbers, the drag coefficient stays.
  89. [89]
    [PDF] Dimensional analysis and scaling laws - Galileo
    We can “solve” certain physical problems with- out actually doing the detailed calculations needed for a complete solution.
  90. [90]
    [PDF] 4 The Behavior of a Simple Pendulum and a Precision ...
    While dimensional analysis cannot determine the value of the number C, it correctly predicts the form of the dependence of the period on quantities that have ...
  91. [91]
    None
    Summary of each segment:
  92. [92]
    [PDF] 8.033 (F24): Lecture 08: Using 4-Momentum - MIT OpenCourseWare
    E = γ(v)mc 2 , p = γ(v)mv ,. (8.26) where v is that body's 3-velocity. This tells us that if we know a body's relativistic energy and relativistic momentum ...
  93. [93]
    [PDF] RELATIVISTIC ENERGY AND MOMENTUM - UT Physics
    momentum p = γmv. Thus, the action S = −mc2. Z worldline dτ,. (90) the Lagrangian L = −mc2q1 − v2/c2,. (91) the momentum p = +mγv = mv p1 − (v/c)2 . (92).
  94. [94]
    [PDF] The proper-force 3-vector - UMSL
    The distinction between proper (i.e. cell-phone detectable) and geometric (i.e. connection-coefficient) forces allows one to use Newton's 3-vector laws in ...
  95. [95]
    [PDF] Special Relativity
    Oct 6, 2010 · The 4-force can be written in terms of the 3-force as. . F. f = d p/dτ = (d p/dt)(dt/dτ) = γ F. ☛ 4-force in terms of 3-force where is the ...Missing: four- limit
  96. [96]
    [PDF] Lecture 1b - U.S. Particle Accelerator School
    Example: Consider a light bulb hanging in a boxcar moving at relativistic velocity. How long does it take a light ray, moving directly down in the boxcar ...
  97. [97]
    28.5 Relativistic Momentum – College Physics - UCF Pressbooks
    Relativistic momentum is classical momentum multiplied by a relativistic factor, using rest mass and velocity relative to an observer.
  98. [98]
    General Relativity - University of Pittsburgh
    In relativity, both special and general, they are timelike geodesics. That is, they are curves of greatest proper time, which is the analog of the straight ...
  99. [99]
    4. Gravitation - Lecture Notes on General Relativity - S. Carroll
    We are now prepared to examine the physics of gravitation as described by general relativity. This subject falls naturally into two pieces.
  100. [100]
    [PDF] Quantum Field Theory - DAMTP
    where the field strength is defined by. Fµν = ∂µAν − ∂νAµ. (6.2). The equations of motion which follow from this Lagrangian are. ∂µ. ∂L. ∂(∂µAν). = −∂µFµν ...
  101. [101]
    [1503.05007] The Evolution of Quantum Field Theory, From QED to ...
    Mar 17, 2015 · Abstract page for arXiv paper 1503.05007: The Evolution of Quantum Field Theory, From QED to Grand Unification. ... Standard Model. Crucial for ...
  102. [102]
    [PDF] 2. The Lagrangian Formalism - DAMTP
    The equations of motion arising from L are m3 x = 2m3! ˙y + m3!2x. @V. @x m3 = 2m3! ˙x + m3!2y. @V. @y. (2.94). The full solutions to these equations are ...
  103. [103]
    Lagrangian formalism for fields - Scholarpedia
    Aug 30, 2010 · The Lagrangian formalism is one of the main tools of the description of the dynamics of a vast variety of physical systems.
  104. [104]
    Conservation of Isotopic Spin and Isotopic Gauge Invariance
    The paper explores local isotopic spin rotations, leading to isotopic gauge invariance and a b field related to isotopic spin, similar to the electromagnetic ...
  105. [105]
    [PDF] quantum yang–Mills Theory - Clay Mathematics Institute
    The non-abelian gauge theory of the strong force is called. Quantum Chromodynamics (QCD). The use of QCD to describe the strong force was motivated by a ...
  106. [106]
    [PDF] Quantum field theory and the Standard Model
    We give an introduction to the elements of quantum field theory, to the construction of the basic Lagrangian for a general gauge theory, and pro- ceed with the ...
  107. [107]
    [PDF] Advanced information on the Nobel Prize in Physics, 5 October 2004
    Oct 5, 2004 · The failure of the Yukawa field theory to describe the strong nuclear forces satisfactorily led the physics community to doubt the relevance of ...
  108. [108]
    Four Fundamental Interaction
    The residual strong force between two protons can be described by the exchange of a neutral pion. Note, the W± is not included as an exchange particle for the ...
  109. [109]
    None
    Nothing is retrieved...<|separator|>
  110. [110]
    DOE Explains...The Weak Force - Department of Energy
    In the Standard Model of particle physics, the weak force is carried by subatomic particles called W and Z bosons. By changing quarks to other types of quarks, ...
  111. [111]
    Experimental Test of Parity Conservation in Beta Decay | Phys. Rev.
    Experimental test of parity conservation in beta decay. CS Wu, E. Ambler, RW Hayward, DD Hoppes, and RP Hudson. Columbia University, New York, New York.
  112. [112]
    Gargamelle - CERN
    The signature of a neutral current event was an isolated vertex from which only hadrons were produced. By July 1973 they had confirmed as many as 166 ...
  113. [113]
    The Standard Model | CERN
    The Standard Model explains how the basic building blocks of matter interact, governed by four fundamental forces.
  114. [114]
    History of dark matter | Rev. Mod. Phys.
    The standard model of modern cosmology is unthinkable without dark matter, although direct detections are still missing. A broad perspective of how dark ...