Fact-checked by Grok 2 weeks ago

Equations of motion

The equations of motion are equations in physics that describe the behavior of a in terms of its motion as a function of time. In , a fundamental example is the set of kinematic equations that relate the , , and of an object, particularly under constant acceleration in one dimension. These kinematic equations provide a framework for analyzing without considering the forces causing it, assuming uniform acceleration. They form the foundation of , the branch of focused on describing motion. The standard kinematic equations for constant acceleration are derived from the definitions of average and by integrating over time. The three fundamental equations are: A fourth equation, x = x_0 + \frac{(v + v_0)}{2} t, follows from the definition of average velocity. These apply to scenarios like under , where is (a = g \approx 9.8 \, \mathrm{m/s^2}). More broadly, equations of motion encompass the equations governing dynamic systems, originating from Newton's second law (\mathbf{F} = m \mathbf{a}), which express as the net force divided by . For variable forces, these yield second-order equations solved numerically or analytically for specific cases. In advanced mechanics, formulations like Lagrange's equations (\frac{d}{dt} \left( \frac{\partial L}{\partial \dot{q}} \right) - \frac{\partial L}{\partial q} = 0, where L is the ) or Hamilton's equations provide generalized equations of motion for complex systems, including constraints and multiple . These equations extend to two- and three-dimensional motion by applying them separately to each component, enabling analysis of trajectories and when combined with vector decompositions. They underpin applications in , , and , from predicting orbits to designing .

Historical Development

Early Conceptual Foundations

The foundational concepts of motion in Western philosophy originated with Aristotle (384–322 BCE), who distinguished between natural and violent motion in his works Physics and On the Heavens. Natural motion occurs when bodies seek their proper place in the cosmos without external interference: heavy elements like earth and water move rectilinearly downward toward the Earth's center, light elements like air and fire move upward, while celestial bodies, composed of ether, execute perfect circular motion around the center. In contrast, violent motion—such as a thrown projectile—deviates from this natural tendency and requires a continuous external force to sustain it, as the body naturally resists such displacement and seeks to return to rest in its natural position. Aristotle categorized motions primarily as rectilinear for sublunary terrestrial bodies and circular for the eternal, unchanging heavens, reflecting his teleological view that all change serves an inherent purpose. Aristotle's qualitative "laws" of motion lacked mathematical precision, relying instead on observational principles; for instance, he posited that heavier objects fall faster than lighter ones because their greater imparts a stronger natural tendency toward the center, with speed proportional to weight. This framework dominated European thought for over a , emphasizing motion as a process toward rather than a quantifiable dynamic. Medieval scholars began critiquing and refining Aristotelian ideas, particularly through the development of impetus theory, which addressed inconsistencies in explaining sustained without perpetual force. Jean Buridan (c. 1300–1361), a French philosopher at the , introduced impetus as an internalized motive quality imparted to a body by the initial mover, allowing it to continue moving after the force ceases, much like a precursor to the concept of . Buridan applied this to both terrestrial projectiles and rotation, suggesting that God initially gave the heavens an impetus that persists eternally due to the lack of resisting medium. Building on Buridan's work, Nicole Oresme (c. 1320–1382), a French theologian and mathematician, further elaborated impetus theory by incorporating graphical representations of velocity and acceleration, demonstrating how motion could vary uniformly or difformly without constant external causes. Oresme used these ideas to challenge Aristotelian uniform speed in falling bodies, proposing that impetus could accumulate, leading to acceleration, though still within a qualitative, non-mathematical framework. These medieval advancements laid intuitive groundwork for later quantitative treatments, influencing thinkers like Galileo in the 16th century who built upon impetus to conduct empirical experiments on motion.

Galilean and Newtonian Advances

advanced the understanding of motion through experimental investigations in the early 17th century, particularly via his experiments, which demonstrated that objects undergo uniform . By rolling balls down smoothly grooved inclines of varying angles, he measured the distances traveled over equal time intervals and found that the distance was proportional to the square of the time, implying a constant acceleration g independent of the object's . These experiments corrected the Aristotelian notion that falling bodies accelerate proportionally to their weight, showing instead that all objects fall with the same acceleration in the absence of air resistance. In his seminal work Dialogues Concerning Two New Sciences, published in 1638, Galileo synthesized these findings and extended them to projectile motion, proving that the trajectory of a projectile in a vacuum follows a parabolic path. This resulted from combining uniform horizontal motion with vertically accelerated free fall under constant gravity. He derived the kinematic equation for free fall from geometric considerations of velocity-time relationships, where the average velocity is half the final velocity, yielding the distance s traveled as s = \frac{1}{2} g t^2 for an object starting from rest, with g as the constant gravitational acceleration. Building on such kinematic insights and Johannes Kepler's empirical laws of planetary motion—which described elliptical orbits with periods squared proportional to semi-major axes cubed—Isaac Newton provided a dynamical framework in his Philosophiæ Naturalis Principia Mathematica (1687). Newton synthesized the concepts of inertia (from Galileo's work), impressed forces, and acceleration into his three laws of motion, with the second law stating that the change in motion (momentum) is proportional to the motive force and occurs in the direction of that force, mathematically expressed as F = ma, where F is the net force, m the mass, and a the acceleration. This equation became the core dynamic relation for equations of motion, enabling Newton to derive Kepler's laws from his universal law of gravitation as a special case.

Post-Newtonian Refinements

Building upon the Newtonian foundations of the late , 18th- and 19th-century mathematicians and physicists extended the equations of motion through more abstract and general frameworks, emphasizing variational principles and coordinate independence. Leonhard Euler made significant contributions to in the mid-18th century, developing equations that describe the rotational motion of solid bodies using vector formulations. In his 1765 work Theoria motus corporum solidorum, Euler derived the equations governing the of rigid bodies, introducing the concept of the inertia tensor and establishing the Euler equations for free rotation, which express the time evolution of components in the body frame. These advancements provided a systematic treatment of three-dimensional rotations, moving beyond particle mechanics to encompass extended objects. Euler also first formulated the differential equation now known as the Euler-Lagrange equation in the 1750s, in connection with problems in the . A key precursor to these developments was Jean le Rond of , formulated around 1743–1750, which reformulates Newton's laws for systems with constraints by considering displacements that satisfy kinematic restrictions. D'Alembert's approach equates the virtual work of applied forces to the negative of the virtual work of inertial forces, enabling the analysis of without explicitly resolving constraint forces. This principle laid the groundwork for more elegant derivations of equations of motion in complex systems. Joseph-Louis Lagrange further generalized these ideas in his 1788 treatise Mécanique Analytique, shifting the focus from force-based descriptions to a coordinate-free formulation using that inherently account for constraints. Lagrange's method avoids direct computation of forces, instead deriving equations of motion from a scalar function—the , defined as kinetic minus —applied to arbitrary systems of particles or rigid bodies. This analytical approach unified diverse mechanical problems under a single framework, emphasizing variational principles over geometric constructions. Lagrange derived the equations by extending to dynamics, applying variations to the in . Central to Lagrange's formulation is the Euler-Lagrange equation, which governs the evolution of q_i and their velocities \dot{q}_i: \frac{d}{dt} \left( \frac{\partial L}{\partial \dot{q}_i} \right) - \frac{\partial L}{\partial q_i} = 0 Later, in 1834, provided a variational derivation of this equation using the principle of stationary , where the integral S = \int_{t_1}^{t_2} L(q, \dot{q}, t) \, dt is made stationary. For systems with conservative forces, where the potential depends only on positions, the is time-independent, leading to along the system's .

Kinematic Equations

Constant Acceleration in Straight-Line Motion

In kinematics, the equations of motion for constant acceleration describe the relationship between , , time, and in one-dimensional straight-line motion. s(t) represents the of an object from a reference point at time t, v(t) = \frac{ds}{dt} is the rate of change of , and a = \frac{dv}{dt} = \frac{d^2s}{dt^2} is the constant rate of change of ./Book:University_Physics_I-Mechanics_Sound_Oscillations_and_Waves(OpenStax)/02%3A_Motion_Along_a_Straight_Line/2.04%3A_Position_Velocity_and_Acceleration) These equations, often referred to as the SUVAT equations (from the variables displacement s, initial velocity u, final velocity v, acceleration a, and time t), are derived directly from the definitions of velocity and acceleration under the assumption of constant a. Integrating acceleration with respect to time yields velocity as v = u + at, where u is the initial velocity at t = 0. Further integration gives position as s = ut + \frac{1}{2}at^2. Eliminating time from these relations produces the equation v^2 = u^2 + 2as. These derivations assume motion in an inertial reference frame, where acceleration remains uniform without varying external influences such as friction or non-constant forces./Book:University_Physics_I-Mechanics_Sound_Oscillations_and_Waves(OpenStax)/02%3A_Motion_Along_a_Straight_Line/2.07%3A_Falling_Objects) A classic example is free fall under gravity near Earth's surface, where acceleration a = g \approx 9.80665 \, \mathrm{m/s^2} acts downward. For an object dropped from rest (u = 0), the position after time t is s = \frac{1}{2}gt^2, and velocity is v = gt. Consider a ball released from a height of 20 m: after 2 seconds, its velocity is v = 9.80665 \times 2 \approx 19.61 \, \mathrm{m/s}, and displacement is s = 0 + \frac{1}{2} \times 9.80665 \times 4 \approx 19.61 \, \mathrm{m}, nearly reaching the ground. Graphically, constant appears as a straight line on a velocity-time plot, with equal to a and area under the curve giving s. On a position-time , the curve is parabolic, reflecting the quadratic dependence on time. These visualizations, originating from Galileo's inclined plane experiments, aid in understanding the uniformity of .

Acceleration in Planar and Spatial Trajectories

In planar and spatial trajectories, the equations of motion for a particle under constant acceleration are expressed in vector form to describe motion in two or three dimensions. The position vector \mathbf{r}(t) of the particle at time t is related to the initial position \mathbf{r}_0, initial velocity \mathbf{v}_0, and constant acceleration vector \mathbf{a} by the equation \mathbf{r}(t) = \mathbf{r}_0 + \mathbf{v}_0 t + \frac{1}{2} \mathbf{a} t^2, while the \mathbf{v}(t) is given by \mathbf{v}(t) = \frac{d\mathbf{r}}{dt} = \mathbf{v}_0 + \mathbf{a} t. These equations generalize the one-dimensional kinematic relations to non-collinear paths, where remains constant in magnitude and direction. A key feature of these equations is the of motion into components along directions, such as Cartesian coordinates. For instance, in three dimensions, the \mathbf{a} = a_x \mathbf{i} + a_y \mathbf{j} + a_z \mathbf{k} leads to separate scalar equations for each component: x(t) = x_0 + v_{0x} t + \frac{1}{2} a_x t^2, and similarly for y(t) and z(t). This reveals that motion in directions to the proceeds with uniform , as the component of \mathbf{a} in those directions is zero. A classic application is projectile motion in a plane, where gravity provides a constant acceleration \mathbf{a} = -g \mathbf{k} directed downward, with g \approx 9.8 \, \mathrm{m/s^2}. Assuming launch from the origin with initial velocity \mathbf{v}_0 = v_0 \cos\theta \, \mathbf{i} + v_0 \sin\theta \, \mathbf{j}, the trajectory is parabolic, and the horizontal range R over level ground is R = \frac{v_0^2 \sin(2\theta)}{g}, maximized at \theta = 45^\circ. This kinematic description applies to ballistic trajectories, such as those of thrown objects or unpowered projectiles, focusing solely on position and velocity evolution without considering force origins.

General Descriptions of Particle Motion

In the general case of particle motion in two or three dimensions, where acceleration \mathbf{a}(t) is an arbitrary function of time and not necessarily constant, the velocity \mathbf{v}(t) is obtained by integrating the acceleration with respect to time, yielding \mathbf{v}(t) = \mathbf{v}_0 + \int_0^t \mathbf{a}(\tau) \, d\tau, where \mathbf{v}_0 is the initial velocity. Similarly, the position \mathbf{r}(t) follows from integrating the velocity, giving \mathbf{r}(t) = \mathbf{r}_0 + \int_0^t \mathbf{v}(\tau) \, d\tau, with \mathbf{r}_0 as the initial position; these relations hold in and describe the kinematic evolution without reference to underlying forces. This integral formulation generalizes the simpler algebraic equations that apply under constant acceleration, reducing to them when \mathbf{a}(t) is time-independent. To characterize the of the particle's in three dimensions, concepts from such as \kappa and torsion \tau provide essential descriptions, independent of the parametrization by time. The Frenet-Serret formulas relate the derivatives of the unit \mathbf{T}, \mathbf{N}, and binormal \mathbf{B} vectors along the curve, parametrized by s, as follows: \frac{d\mathbf{T}}{ds} = \kappa \mathbf{N}, \quad \frac{d\mathbf{N}}{ds} = -\kappa \mathbf{T} + \tau \mathbf{B}, \quad \frac{d\mathbf{B}}{ds} = -\tau \mathbf{N}. These equations describe how the twists and bends, with the speed v = ds/dt linking the time parametrization to the spatial curve; measures the instantaneous rate of turning, while torsion quantifies the out-of-plane deviation. In , they apply to any smooth particle path, enabling analysis of the 's intrinsic properties. A key aspect of such general motion is its representation in , where the state of the particle is depicted by the pair (\mathbf{r}(t), \mathbf{v}(t)) evolving along a in a $2n-dimensional (for n spatial dimensions); this highlights the deterministic flow governed by the kinematic integrals but reveals no closed-form solutions unless \mathbf{a}(t) is explicitly specified, often requiring numerical approximation for complex cases. For instance, in non-uniform gravitational fields where varies spatially and temporally, such as near extended masses, the must typically be computed numerically using methods like Euler integration, which approximates the solution by stepping forward in time via \mathbf{v}(t + \Delta t) \approx \mathbf{v}(t) + \mathbf{a}(t) \Delta t and \mathbf{r}(t + \Delta t) \approx \mathbf{r}(t) + \mathbf{v}(t) \Delta t, though higher-order schemes improve accuracy for stiff problems. Beyond acceleration, higher-order kinematic quantities extend the description; the jerk \mathbf{j}(t) = d\mathbf{a}/dt represents the time rate of change of acceleration, capturing abrupt variations in motion that affect smoothness and comfort in applications like vehicle dynamics. This third derivative of position provides insight into the non-linearity of trajectories under variable acceleration, with its magnitude influencing higher-frequency components of the path.

Newtonian Dynamic Equations

Fundamental Laws and Derivations

, first systematically presented by in his 1687 work , provide the foundational principles for . The first law, known as the law of inertia, states that a body remains at rest or in uniform straight-line motion unless acted upon by an external . This law defines the concept of inertial motion and implies the existence of inertial reference frames where no net results in zero . The third law asserts that for every action, there is an equal and opposite reaction, meaning forces between interacting bodies are mutual and collinear but oppositely directed. At the core of dynamic equations is Newton's second law, which relates the net external \mathbf{F} on a body to the rate of change of its linear momentum \mathbf{p}: \mathbf{F} = \frac{d\mathbf{p}}{dt}. Linear momentum is defined as \mathbf{p} = m \mathbf{v}, where m is the and \mathbf{v} is the . For systems of constant mass, this simplifies to \mathbf{F} = m \mathbf{a}, where \mathbf{a} is the . This vector equation governs the motion of particles in inertial frames, where inertial frames are those unaccelerated relative to absolute space, approximately realized by non-accelerating laboratory frames on . To derive equations of motion, apply the second law in an . For a constant , \mathbf{a} is constant since m is typically invariant. Integrating \mathbf{a} = dv/dt yields \mathbf{v} = \mathbf{a} t + \mathbf{v}_0, and further integration gives position \mathbf{r} = \frac{1}{2} \mathbf{a} t^2 + \mathbf{v}_0 t + \mathbf{r}_0. These relations constitute the kinematic equations for constant acceleration, linking forces directly to trajectories. In non-inertial frames, such as rotating or accelerating ones, Newton's laws require modification; fictitious forces emerge to account for the frame's motion, including the (directed outward from the rotation axis) and the (proportional to cross , deflecting moving objects). A representative application is the spring-block system, where a m is attached to a with constant k on a frictionless surface. The restoring force follows , \mathbf{F} = -k \mathbf{x}, where \mathbf{x} is displacement from equilibrium. Substituting into the second law gives m \frac{d^2 \mathbf{x}}{dt^2} = -k \mathbf{x}, or \frac{d^2 \mathbf{x}}{dt^2} + \frac{k}{m} \mathbf{x} = 0. This second-order describes , with solutions of the form \mathbf{x}(t) = A \cos(\omega t + \phi), where \omega = \sqrt{k/m} is the .

Applications to Rigid Bodies and Systems

In Newtonian mechanics, the translational motion of a is governed by the equation for its , where the net external force \mathbf{F} equals the total M times the acceleration of the \mathbf{a}_\text{cm}, \mathbf{F} = M \mathbf{a}_\text{cm}. This equation extends the single-particle form \mathbf{F} = m \mathbf{a} to extended bodies by treating the as an effective particle under the influence of all external forces. The rotational of a about its involve the net external \boldsymbol{\tau}, which equals the time derivative of the \mathbf{L}, \boldsymbol{\tau} = \frac{d\mathbf{L}}{dt}. For a , \mathbf{L} = \mathbf{I} \boldsymbol{\omega}, where \mathbf{I} is the tensor and \boldsymbol{\omega} is the vector. The tensor \mathbf{I} is a symmetric 3×3 matrix that quantifies the relative to the chosen axes, with diagonal elements as principal moments of inertia and off-diagonal elements as products of inertia. In the body frame aligned with the principal axes, where \mathbf{I} is diagonal, Euler's equations simplify the rotational to: I_1 \dot{\omega}_1 + (I_3 - I_2) \omega_2 \omega_3 = \tau_1, I_2 \dot{\omega}_2 + (I_1 - I_3) \omega_3 \omega_1 = \tau_2, I_3 \dot{\omega}_3 + (I_2 - I_1) \omega_1 \omega_2 = \tau_3, describing the evolution of \boldsymbol{\omega} under applied torques \boldsymbol{\tau}. A classic example is the compound pendulum, an extended rigid body pivoting about a fixed axis not through its center of mass, such as a uniform rod of length L and mass m suspended from one end. The torque due to gravity about the pivot yields the equation I \ddot{\theta} = -mg \frac{L}{2} \sin \theta, where I = \frac{1}{3} m L^2 is the moment of inertia about the pivot and \theta is the angular displacement from vertical, leading to small-angle oscillatory motion with period T = 2\pi \sqrt{\frac{2L}{3g}}. Another illustrative case is rolling without slipping down an incline, where for a body of mass m, radius r, and moment of inertia I about the center, the linear acceleration a satisfies a = \frac{g \sin \theta}{1 + \frac{I}{m r^2}}, combining translational and rotational equations with the no-slip condition a = r \alpha. For a solid sphere, I = \frac{2}{5} m r^2, yielding a = \frac{5}{7} g \sin \theta. For systems of interacting rigid bodies, constraints such as joints or contacts introduce reaction forces that must be accounted for without altering the underlying Newtonian equations. In the Newtonian framework, incorporates these via , and Lagrange multipliers \lambda_k can be introduced to enforce m g_k(\mathbf{q}, t) = 0, modifying the equations to \mathbf{F} - \sum_k \lambda_k \frac{\partial g_k}{\partial \mathbf{q}} = m \ddot{\mathbf{q}} for coordinates \mathbf{q}. This approach efficiently solves for both motion and constraint forces in multi-body systems like linkages. A key application is the central force problem in two-body systems, where the interaction depends only on separation distance, reducing the motion to an effective one-body problem in a with conserved , yielding conic-section orbits such as ellipses for inverse-square forces like . Conservation laws in these systems arise from spatial symmetries: translation invariance implies conservation of total linear momentum \mathbf{P} = \sum M_i \mathbf{v}_{\text{cm},i} if no external forces act, while rotational invariance about a point conserves total angular momentum \mathbf{L} = \sum (\mathbf{r}_i \times M_i \mathbf{v}_{\text{cm},i} + \mathbf{L}_i), where \mathbf{L}_i is each body's spin angular momentum. These principles, derived from in the context but verifiable directly from Newton's laws, underpin the stability of and multi-body dynamics.

Formulations in Analytical Mechanics

Lagrangian Approach

The Lagrangian approach to deriving equations of motion reformulates in terms of functionals and variational principles, providing a systematic method to obtain the dynamics of systems with multiple . The L is defined as the difference between the T and the V of the system, L = T - V. This formulation, introduced by in his seminal work Mécanique Analytique, allows for the use of q_i that may not correspond directly to Cartesian positions, making it particularly suited for systems with constraints or complex geometries. The equations of motion arise from the principle of stationary , which states that the physical path of the system extremizes the integral S = \int_{t_1}^{t_2} L(q_i, \dot{q}_i, t) \, dt. Varying this integral and setting the first variation \delta S = 0 yields the Euler-Lagrange equations for each generalized coordinate q_i: \frac{d}{dt} \left( \frac{\partial L}{\partial \dot{q}_i} \right) - \frac{\partial L}{\partial q_i} = 0. These second-order differential equations describe the system's evolution without explicitly invoking forces, as in Newtonian . For conservative systems, this approach is equivalent to Newton's second , but it naturally incorporates (those expressible as functions of coordinates, such as fixed lengths in pendulums) by reducing the number of independent coordinates, avoiding the need for Lagrange multipliers in many cases. A classic example is the simple pendulum, where the generalized coordinate is the angle \theta from the vertical, with length l and mass m. The kinetic energy is T = \frac{1}{2} m l^2 \dot{\theta}^2 and the potential energy is V = -m g l \cos \theta, yielding L = \frac{1}{2} m l^2 \dot{\theta}^2 + m g l \cos \theta. Applying the Euler-Lagrange equation gives the nonlinear equation of motion \ddot{\theta} + \frac{g}{l} \sin \theta = 0, which for small angles approximates simple harmonic motion. For more complex systems, such as the double pendulum with two masses m_1, m_2 and lengths l_1, l_2, the Lagrangian involves coupled angles \theta_1, \theta_2, leading to a set of four first-order equations that exhibit chaotic behavior for sufficiently large initial displacements due to the nonlinearity. Key advantages of the Lagrangian formulation include its coordinate independence, which simplifies calculations in non-Cartesian systems like , and its facilitation of symmetry analysis. Notably, establishes that every continuous symmetry of the action corresponds to a : if the Lagrangian is invariant under a transformation \delta q_i = \epsilon K_i(q, t), then the quantity \sum_i \frac{\partial L}{\partial \dot{q}_i} K_i is conserved along the system's trajectory. For instance, time-translation invariance implies , while spatial translation invariance yields conservation, providing a deep link between symmetries and the integrals of motion. This theorem, proven by in her 1918 paper, underpins much of modern .

Hamiltonian Approach

The Hamiltonian approach provides a reformulation of in , consisting of q_i and their conjugate momenta p_i, emphasizing the symplectic structure of the dynamics. This framework, developed by in his 1834 paper "On a General Method in Dynamics," shifts the focus from velocities to momenta, facilitating the analysis of and long-term behavior in conservative systems. The approach derives from the formulation through a , where the momenta are defined as p_i = \frac{\partial L}{\partial \dot{q}_i} and the H(q, p, t) = \sum_i p_i \dot{q}_i - L(q, \dot{q}, t), with velocities \dot{q}_i expressed as functions of q and p. For scleronomic systems without explicit time dependence in the , the represents the total energy, H = T + V, where T is the and V is the , both rewritten in terms of q and p. The equations of motion emerge as Hamilton's canonical equations: \dot{q}_i = \frac{\partial H}{\partial p_i}, \quad \dot{p}_i = -\frac{\partial H}{\partial q_i}. These differential equations generate the trajectories in and are equivalent to the second-order Euler-Lagrange equations but offer greater symmetry and utility for transformations. A key feature of the Hamiltonian formalism is the existence of canonical transformations, which map old coordinates and momenta (q, p) to new ones (Q, P) while preserving the form of Hamilton's equations, provided the transformation satisfies \sum_i (p_i dq_i - P_i dQ_i) = dF for some F. Such transformations simplify complex problems, like reducing the number of in integrable systems. Another fundamental property is , which states that the phase-space volume occupied by an ensemble of trajectories remains constant over time, implying incompressible flow in and conservation of information under Hamiltonian evolution. Illustrative examples highlight the approach's power. For the one-dimensional , the is H = \frac{p^2}{2m} + \frac{1}{2} k q^2, yielding \dot{q} = \frac{p}{m} and \dot{p} = -k q, whose solutions describe periodic motion with constant energy. In central problems using polar coordinates (r, \theta), the becomes H = \frac{p_r^2}{2m} + \frac{p_\theta^2}{2m r^2} + V(r), where p_\theta is conserved (), allowing separation of radial and angular equations for bound orbits like Keplerian ellipses. The Hamiltonian formulation excels in connecting to , as the conserved phase volume from underpins the and for equilibrium distributions. It also provides criteria for integrability, such as the existence of action-angle variables through canonical transformations, enabling quasi-periodic solutions via frequency analysis.

Equations in Electrodynamics

Motion of Charged Particles

The motion of charged particles in electromagnetic fields is described by the law, which determines the experienced by a particle due to the interaction with and magnetic fields. This force arises from the fundamental principles of and serves as the basis for the equations of motion in electrodynamics. For a particle of charge q and \mathbf{v}, the \mathbf{F} is given by \mathbf{F} = q \left( \mathbf{E} + \mathbf{v} \times \mathbf{B} \right), where \mathbf{E} is the electric field and \mathbf{B} is the magnetic field. This expression, originally developed by Hendrik Lorentz in his 1895 treatise on electromagnetic phenomena in moving bodies, combines the Coulomb force from the electric field and the magnetic force perpendicular to both the velocity and the magnetic field. The resulting equation of motion in the non-relativistic limit is Newton's second law adapted to electromagnetic forces: m \mathbf{a} = q \left( \mathbf{E} + \mathbf{v} \times \mathbf{B} \right), where m is the particle mass and \mathbf{a} = d\mathbf{v}/dt is the acceleration. This formulation assumes constant fields and neglects higher-order effects like radiation reaction, focusing solely on the direct field-particle interaction. In uniform constant fields, the trajectories can be solved analytically, revealing characteristic behaviors. For a pure (\mathbf{E} = 0), the magnetic force provides the centripetal acceleration for in the plane perpendicular to \mathbf{B}, with the cyclotron frequency \omega_c = q B / m dictating the angular speed of gyration. The radius of this orbit, known as the , is r = m v_\perp / (q B), where v_\perp is the component perpendicular to \mathbf{B}. If the initial has a component parallel to \mathbf{B}, the path becomes helical, combining uniform motion along the field lines with circular gyration. A representative example is an in a uniform , such as in Earth's or laboratory devices, where it traces helical paths with gyroradii on the order of micrometers for typical field strengths of 1 and thermal velocities (e.g., at 1 ). When both electric and magnetic fields are present and uniform, the motion includes drift effects superimposed on the orbits. In the , for instance, charged carriers in a current-carrying conductor experience a magnetic force perpendicular to their , deflecting them to one side of the material and creating a transverse (Hall field) that balances the magnetic force in . This phenomenon, first observed by Edwin Hall in 1879, provides a measure of carrier density and type (electrons or holes) through the Hall voltage V_H = I B / (n e d), where I is current, n is carrier density, e is , and d is thickness. For metals like silver ribbons with currents around 1 A and fields of 0.1 T, Hall voltages reach microvolts, illustrating the scale of deflection for conduction electrons. For high-speed particles where velocities approach the c, relativistic effects modify the dynamics. The relativistic equation of motion is \frac{d}{dt} (\gamma m \mathbf{v}) = q \left( \mathbf{E} + \mathbf{v} \times \mathbf{B} \right), where \gamma = (1 - v^2/c^2)^{-1/2} is the . This form ensures consistency with , altering the effective inertia and leading to velocity-dependent trajectories, such as compressed gyroradii in strong fields for particles like cosmic-ray protons. Detailed relativistic treatments, including formulations, are covered in the context of special relativistic dynamics.

Coupled Field-Particle Dynamics

In coupled field-particle dynamics, the equations of motion for charged particles are intertwined with the evolution of electromagnetic fields, forming a self-consistent system where particle currents and charges source the fields, while the fields exert forces on the particles. This framework extends beyond the test-particle approximation by accounting for the feedback effects of particle motion on field generation, essential for describing collective phenomena in systems like plasmas and high-energy accelerators. The core coupling arises through , where the \mathbf{J} and \rho are expressed as sums over particles: \rho(\mathbf{r}, t) = \sum_i q_i \delta(\mathbf{r} - \mathbf{r}_i(t)) and \mathbf{J}(\mathbf{r}, t) = \sum_i q_i \mathbf{v}_i(t) \delta(\mathbf{r} - \mathbf{r}_i(t)), driving the field dynamics via \nabla \cdot \mathbf{E} = \rho / \epsilon_0 and \nabla \times \mathbf{B} = \mu_0 \mathbf{J} + \mu_0 \epsilon_0 \partial_t \mathbf{E} (in SI units). For a single accelerated charge, the self-interaction with its own introduces a radiation reaction force, captured by the Abraham-Lorentz in the non-relativistic : m \dot{\mathbf{v}} = \mathbf{F}_\text{ext} + \frac{2 q^2}{3 c^3} \ddot{\mathbf{v}}, where the additional term proportional to \ddot{\mathbf{v}} represents energy loss to radiated electromagnetic waves. This equation, derived from the conservation of energy-momentum in the , highlights the back-reaction but suffers from issues like runaway solutions, often mitigated by approximations such as the Landau-Lifshitz reduction that avoids higher derivatives. In relativistic extensions, the Lorentz-Dirac equation generalizes this form covariantly, incorporating effects crucial for ultra-relativistic particles. In multi-particle systems, such as collisionless plasmas, the Vlasov-Maxwell system provides the statistical description: the \frac{\partial f}{\partial t} + \mathbf{v} \cdot \nabla f + \frac{q}{m} (\mathbf{E} + \mathbf{v} \times \mathbf{B}) \cdot \nabla_v f = 0 governs the phase-space distribution f(\mathbf{r}, \mathbf{v}, t), self-consistently coupled to via moments \rho = q \int f \, d^3\mathbf{v} and \mathbf{J} = q \int \mathbf{v} f \, d^3\mathbf{v}. Originally proposed for kinetic theory in conducting fluids, this set enables analysis of collective excitations like Langmuir waves and instabilities, where particle distributions evolve under mean self-fields. For fluid-like behaviors in dense plasmas, (MHD) approximates the Vlasov system by taking velocity moments, yielding coupled equations for bulk flow \mathbf{v}, density \rho, and fields, such as the induction equation \partial_t \mathbf{B} = \nabla \times (\mathbf{v} \times \mathbf{B} - \eta \nabla \times \mathbf{B}) under ideal conductivity \eta \to 0, describing phenomena like in astrophysical contexts. Applications abound in high-energy physics, notably particle accelerators where relativistic electrons in curved trajectories emit , necessitating inclusion of radiation reaction in the equations of motion to model loss and ; for instance, in rings, the power radiated scales as \propto \gamma^4 / \rho (with \gamma and bending radius \rho), influencing design limits for facilities like the LHC. In contexts, MHD approximations simplify to resistive or ideal forms for devices and flares. Numerical resolution of these coupled systems relies heavily on particle-in-cell (PIC) simulations, which track superparticles on a to compute self-consistent fields via finite-difference solutions to , enabling studies of wave-particle interactions with high fidelity in three dimensions. This method, foundational since the 1960s, remains central to 2025-era computations for laser- acceleration and modeling.

Relativistic Equations of Motion

Special Relativistic Kinematics and Dynamics

In , kinematics describes the motion of objects at speeds approaching the c, where classical Newtonian concepts fail due to the absence of an absolute . This framework, developed by , posits that all inertial frames are equivalent, with the laws of physics invariant under Lorentz transformations. \tau serves as the invariant interval along a particle's worldline, defined as d\tau = dt / \gamma, where t is coordinate time and \gamma = 1 / \sqrt{1 - v^2/c^2} is the , with v the particle's speed. The u^\mu = dx^\mu / d\tau is a tangent to the worldline, satisfying u^\mu u_\mu = -c^2 (in the mostly-plus metric convention), ensuring its magnitude is constant at c. This formulation, introduced by , unifies space and time into Minkowski spacetime, enabling covariant descriptions of motion. Velocity addition in special relativity deviates from vector summation to preserve the constancy of c. For collinear velocities u and v along the x-axis, the composed velocity w in the lab frame is w = (u + v) / (1 + uv/c^2), derived from Lorentz transformations applied to velocity components. This formula ensures that if v = c (light), w = c regardless of u, upholding the second postulate of relativity. For non-collinear cases, the general addition uses Lorentz boosts, preventing superluminal speeds. In the low-speed limit v \ll c, it reduces to the classical w \approx u + v. Hyperbolic motion exemplifies uniform proper acceleration \alpha, where the worldline traces a in . The position is given by x = \frac{c^2}{\alpha} \left( \cosh\left(\frac{\alpha \tau}{c}\right) - 1 \right), \quad ct = \frac{c^2}{\alpha} \sinh\left(\frac{\alpha \tau}{c}\right), yielding velocity v = c \tanh(\alpha \tau / c) and \gamma = \cosh(\alpha \tau / c). This motion, first detailed by , models scenarios like rocket propulsion under constant felt acceleration. Relativistic dynamics extends Newton's second law using four-vectors. The four-momentum p^\mu = m u^\mu has spatial part \mathbf{p} = \gamma m \mathbf{v}, where m is rest mass, generalizing classical \mathbf{p} = m \mathbf{v}. The total is E = \gamma m c^2, encompassing rest energy m c^2 and kinetic contributions, as derived by Einstein from thought experiments on energy-momentum conservation. The f^\mu = dp^\mu / d\tau is orthogonal to u^\mu, with three-force \mathbf{F} = d\mathbf{p}/dt = \gamma \mathbf{f}, where \mathbf{f} is the spatial part of f^\mu. For constant proper force, trajectories follow hyperbolic paths analogous to . In particle colliders like the (LHC), these equations govern proton trajectories at v \approx 0.99999999 c, where \gamma \approx 7000, requiring precise Lorentz-invariant beam dynamics to maintain stability and collision efficiency.

Geodesic Motion in General Relativity

In , the trajectory of a subject solely to follows a in the curved manifold, representing the shortest path analogous to a straight line in flat space. This formulation arises from Albert Einstein's , which posits that the effects of are locally indistinguishable from , thereby interpreting gravitational motion as inertial motion along geodesics rather than the action of a force. The equation governs this motion and is derived using the applied to the interval defined by the . The in a general is given by ds^2 = g_{\mu\nu} \, dx^\mu \, dx^\nu, where g_{\mu\nu} is the , and Greek indices run from 0 to 3. For a timelike (corresponding to massive particles), the \tau satisfies ds^2 = -c^2 d\tau^2, and the worldline extremizes the , equivalent to extremizing the S = -mc \int d\tau. Varying this with respect to the coordinates x^\mu(\tau) yields the equation: \frac{d^2 x^\mu}{d\tau^2} + \Gamma^\mu_{\alpha\beta} \frac{dx^\alpha}{d\tau} \frac{dx^\beta}{d\tau} = 0, where \Gamma^\mu_{\alpha\beta} is the affine connection, specifically the Levi-Civita Christoffel symbols for a torsion-free, metric-compatible connection, defined as \Gamma^\mu_{\alpha\beta} = \frac{1}{2} g^{\mu\sigma} \left( \partial_\alpha g_{\beta\sigma} + \partial_\beta g_{\alpha\sigma} - \partial_\sigma g_{\alpha\beta} \right). This equation describes the parallel transport of the tangent vector along the curve, ensuring that the particle experiences no proper acceleration in its local frame. A prominent example is the motion in the , which describes the around a spherically symmetric, non-rotating mass M: ds^2 = -\left(1 - \frac{2GM}{c^2 r}\right) c^2 dt^2 + \left(1 - \frac{2GM}{c^2 r}\right)^{-1} dr^2 + r^2 d\Omega^2. Solving the geodesic equation for timelike paths yields orbital trajectories, including the precession of Mercury's perihelion by approximately 43 arcseconds per century, a prediction first calculated by Einstein that resolved a longstanding discrepancy with Newtonian . For null geodesics (light rays), the equation predicts deflection by massive bodies; Einstein computed a 1.75 arcsecond bending for light grazing the Sun's surface, later confirmed by observations during the 1919 . In the weak-field limit, where gravitational potentials are small (|\Phi| \ll c^2) and velocities are non-relativistic, the metric component simplifies to g_{00} \approx -(1 + 2\Phi/c^2), with spatial components nearly . Substituting into the equation for recovers Newton's second law, d^2 \mathbf{x}/dt^2 = -\nabla \Phi, demonstrating the compatibility of with Newtonian gravity in this regime.

Effects on Spinning Particles

In , the motion of particles with intrinsic deviates from the paths followed by spinless point particles due to the coupling between the spin and . The Mathisson–Papapetrou–Dixon () equations provide the framework for describing this dynamics for test particles, where self-gravitational effects are negligible. These equations extend the equation by incorporating the spin tensor, leading to a more complex worldline that accounts for the finite size and orientation of the particle. The core of the MPD equations is the evolution of the p^\mu along the \tau: \frac{D p^\mu}{d\tau} = -\frac{1}{2} R^\mu{}_{\nu\rho\sigma} u^\nu S^{\rho\sigma}, where \frac{D}{d\tau} denotes the , R^\mu{}_{\nu\rho\sigma} is the , u^\mu is the (normalized as u^\mu u_\mu = -1), and S^{\rho\sigma} is the antisymmetric spin tensor representing the . This force term arises from the spin-curvature coupling, causing the particle's to curve away from a in regions of nonzero curvature. The spin tensor itself evolves via \frac{D S^{\rho\sigma}}{d\tau} = p^{[\rho} u^{\sigma]}, but under the Tulczyjew supplementary condition p_\mu S^{\mu\nu} = 0, which aligns the with the center-of-mass worldline, the spin undergoes Fermi-Walker . This transport law ensures the spin remains constant in magnitude and does not rotate due to the particle's or changes, preserving an intrinsic reference frame for the particle. The MPD equations are typically applied in the pole-dipole approximation, where the particle is modeled with only a () and () moment, neglecting higher multipoles like deformations that would require more detailed stress-energy distributions. This simplification is valid for small, weakly gravitating objects such as neutron stars or fundamental particles treated classically. A prominent example is the of in the , which describes the spacetime around a ; here, the spin-curvature interaction induces Larmor-like precession, altering orbital stability and resonance conditions for test particles. In binary systems, spin effects modify the inspiral and merger phases, enhancing or suppressing emission through and spin-orbit coupling. As of 2025, observations from the LIGO-Virgo-KAGRA collaboration have provided empirical confirmation of these spin effects in astrophysical contexts. For instance, detections of two mergers in late 2024, on October 11 (GW241011) and November 10 (GW241110), announced on October 28, 2025, revealed strongly inclined spin axes and rapid rotations in the merging s, deforming their horizons and influencing the gravitational waveforms in ways consistent with MPD-derived predictions for spinning binaries. These events, involving second-generation s formed from prior mergers, highlight how spin contributes to hierarchical growth and tests the validity of the pole-dipole model at extreme masses.

Analogues in Waves, Fields, and Quantum Theory

Wave Propagation Equations

The classical describes the propagation of waves in various media, such as strings, fluids, and electromagnetic fields, exhibiting motion-like behavior through the spatiotemporal evolution of a disturbance. In one dimension, it takes the form \frac{\partial^2 [\psi](/page/Psi)}{\partial t^2} = c^2 \frac{\partial^2 [\psi](/page/Psi)}{\partial x^2}, where \psi(x,t) represents the wave displacement or , and c is the wave speed. This second-order arises from fundamental physical principles and governs non-dispersive where the speed is independent of . A general to the one-dimensional , known as d'Alembert's solution, is given by \psi(x,t) = f(x - c t) + g(x + c t), where f and g are arbitrary twice-differentiable functions determined by initial conditions. This form illustrates the superposition of two waves traveling in opposite directions at speed c, highlighting the principle of wave propagation without distortion in linear media. The wave equation can be derived from Newton's second law applied to a small element of the medium. For a transverse wave on a taut string under tension T with linear mass density \mu, the net force on a segment of length \Delta x is T \left( \frac{\partial \psi}{\partial x}\bigg|_{x+\Delta x} - \frac{\partial \psi}{\partial x}\bigg|_x \right), leading to \mu \Delta x \frac{\partial^2 \psi}{\partial t^2} = T \frac{\partial^2 \psi}{\partial x^2} \Delta x in the limit \Delta x \to 0, yielding c^2 = T/\mu. Similarly, for longitudinal sound waves in a fluid, the equation emerges from the linearized Euler and continuity equations, relating pressure perturbation p to displacement \xi via the bulk modulus B and density \rho, with \frac{\partial^2 \xi}{\partial t^2} = \frac{B}{\rho} \frac{\partial^2 \xi}{\partial x^2} and c^2 = B/\rho; the pressure satisfies p = -\rho c^2 \frac{\partial \xi}{\partial x}. For electromagnetic waves in vacuum, the wave equation derives from Maxwell's equations: taking the curl of Faraday's law \nabla \times \mathbf{E} = -\frac{\partial \mathbf{B}}{\partial t} and substituting Ampère's law with Maxwell's correction \nabla \times \mathbf{B} = \mu_0 \epsilon_0 \frac{\partial \mathbf{E}}{\partial t} (in source-free vacuum) yields \nabla^2 \mathbf{E} = \frac{1}{c^2} \frac{\partial^2 \mathbf{E}}{\partial t^2}, where c = 1/\sqrt{\mu_0 \epsilon_0}; an analogous equation holds for \mathbf{B}. Plane wave solutions \psi = A e^{i(kx - \omega t)} satisfy the dispersion relation \omega = c k, implying a phase velocity v_p = \omega / k = c and group velocity v_g = d\omega / dk = c, both equal to the wave speed for these non-dispersive systems. Superpositions of such waves form wave packets, where the envelope propagates at the group velocity, analogous to the trajectory of a classical particle localizing the disturbance.

Field Evolution Equations

Field evolution equations govern the temporal and spatial changes of physical fields, extending the concept of equations of motion from point particles to distributed quantities across . These equations typically take the form \partial \phi / \partial t = F[\phi, \nabla \phi], where \phi represents the field configuration, and the functional F incorporates interactions, derivatives, and possibly nonlinear terms that dictate the field's or . This structure parallels Newtonian or relativistic particle dynamics but operates on continuous fields, ensuring laws like energy-momentum through underlying symmetries. A foundational example is the advection equation, \partial \phi / \partial t + \mathbf{v} \cdot \nabla \phi = 0, which describes the of a scalar field \phi by a constant \mathbf{v} without sources or dissipation, commonly arising in for or fields. In relativistic contexts, scalar fields obey the Klein-Gordon equation, (\square - m^2) \phi = 0, where \square = \partial^\mu \partial_\mu is the d'Alembertian and m is the field's mass parameter, capturing wave-like propagation with dispersion for massive particles. This equation originates from the independent works of and Walter Gordon in 1926, who sought a relativistic generalization of the for de Broglie waves. It is derived variationally from the action principle, using the density \mathcal{L} = \frac{1}{2} (\partial^\mu \phi)(\partial_\mu \phi) - \frac{1}{2} m^2 \phi^2, where the Euler-Lagrange equations for fields yield the dynamics, analogous to the formulation for particles. For spin-1/2 fields, the , i \gamma^\mu \partial_\mu \psi - m \psi = 0, provides the relativistic evolution in the classical limit, treating \psi as a classical field before quantization, as formulated by in 1928 to reconcile with . Its derivation similarly proceeds from an action \int \bar{\psi} (i \gamma^\mu \partial_\mu - m) \psi \, d^4x, emphasizing first-order time evolution for . The massless case of the Klein-Gordon equation specializes to the wave equation, describing undamped propagation at the . Key properties of these equations include symmetries that underpin their physical consistency. In electrodynamics, the evolution of the , governed by \partial_\mu F^{\mu\nu} = j^\nu and \partial_{[\lambda} F_{\mu\nu]} = 0, derives from the gauge-invariant \mathcal{L} = -\frac{1}{4} F_{\mu\nu} F^{\mu\nu} - A_\mu j^\mu, where F_{\mu\nu} = \partial_\mu A_\nu - \partial_\nu A_\mu and the four-potential A_\mu transforms as A_\mu \to A_\mu + \partial_\mu \Lambda without altering observables, ensuring redundancy-free descriptions. This gauge invariance, first emphasized in the context of classical field theory, allows flexible choices of gauge while preserving the equations of motion. For fields coupled to gravity, the stress-energy tensor T_{\mu\nu}, which sources curvature in Einstein's field equations G_{\mu\nu} = 8\pi T_{\mu\nu}, emerges as the Noether current from translational invariance of the action, with components like T^{00} representing energy density and T^{0i} momentum flux. In the Standard Model, the Higgs field's dynamics follow a similar Klein-Gordon-like equation with a Mexican-hat potential V(\phi) = -\mu^2 |\phi|^2 + \lambda |\phi|^4, but post-2012 discovery measurements through 2025 have confirmed the metastability of the electroweak vacuum up to high energy scales, with its lifetime far exceeding the age of the universe, although quantum corrections and recent analyses hint at possible deeper minima or new phases at even higher scales.

Quantum Mechanical Motion Operators

In quantum mechanics, the evolution of a system's state is governed by the Schrödinger equation, which describes the time dependence of the wave function \psi(\mathbf{r}, t). This equation, i \hbar \frac{\partial \psi}{\partial t} = \hat{H} \psi, where \hat{H} is the Hamiltonian operator, serves as the quantum analogue to classical equations of motion by dictating how the probability amplitude propagates in configuration space. The Hamiltonian \hat{H} typically includes kinetic and potential energy terms, such as \hat{H} = -\frac{\hbar^2}{2m} \nabla^2 + V(\mathbf{r}) for a single particle, mirroring the classical Hamiltonian but quantized through operators. A complementary perspective arises in the , where operators evolve in time while the state remains fixed. The Heisenberg equation of motion for any operator \hat{A} is given by \frac{d\hat{A}}{dt} = \frac{i}{\hbar} [\hat{H}, \hat{A}] + \frac{\partial \hat{A}}{\partial t}, with the commutator [\hat{H}, \hat{A}] = \hat{H}\hat{A} - \hat{A}\hat{H} driving the dynamics. This formulation parallels classical Poisson brackets via the correspondence principle, where the commutator replaces i\hbar times the bracket. For position \hat{\mathbf{x}} and momentum \hat{\mathbf{p}} operators, it yields \frac{d\hat{\mathbf{x}}}{dt} = \frac{\hat{\mathbf{p}}}{m} and \frac{d\hat{\mathbf{p}}}{dt} = -\left\langle \frac{\partial V}{\partial \mathbf{x}} \right\rangle, assuming a time-independent potential. The connects these quantum equations to classical limits by stating that the expectation values evolve as \frac{d \langle \hat{\mathbf{x}} \rangle}{dt} = \frac{\langle \hat{\mathbf{p}} \rangle}{m} and \frac{d \langle \hat{\mathbf{p}} \rangle}{dt} = -\left\langle \frac{\partial V}{\partial \mathbf{x}} \right\rangle, demonstrating how quantum averages approximate Newtonian motion for macroscopic systems or when wave functions are sharply peaked. This underscores the semiclassical regime where quantum fluctuations are negligible. Illustrative examples highlight these operators in action. For a free particle, the Schrödinger equation admits Gaussian wave packet solutions, where the initial localized packet spreads over time due to dispersion, with width \sigma(t) = \sigma_0 \sqrt{1 + \left(\frac{\hbar t}{2m \sigma_0^2}\right)^2}, reflecting the uncertainty in momentum. In the quantum harmonic oscillator, exact stationary states are Hermite-Gaussian functions, and time evolution under \hat{H} = \frac{\hat{p}^2}{2m} + \frac{1}{2} m \omega^2 \hat{x}^2 produces coherent states that oscillate classically while maintaining minimal uncertainty. Non-classical features emerge through the , \Delta x \Delta p \geq \hbar/2, which imposes fundamental limits on simultaneous measurements of and , altering motion predictability compared to classical . Quantum tunneling exemplifies this, allowing particles to traverse potential barriers forbidden classically; for instance, in , the is \lambda \propto e^{-2\int \sqrt{2m(V-E)} \, [dx](/page/DX) / \hbar}, enabling exponential escape probabilities. Bohmian mechanics offers an alternative interpretation via deterministic trajectories guided by the wave function, where particle positions \mathbf{x}(t) follow \frac{d\mathbf{x}}{dt} = \frac{\hbar}{m} \Im \left( \frac{\nabla \psi}{\psi} \right), recovering quantum statistics through an initial ensemble while providing a non-local, ontological picture of motion. This approach, though controversial for introducing hidden variables, illustrates quantum motion as pilot-wave dynamics.

References

  1. [1]
    Description of Motion - HyperPhysics
    Motion is described in terms of displacement (x), time (t), velocity (v), and acceleration (a). Velocity is the rate of change of displacement.
  2. [2]
    Introduction to Motion – Physics 131 - Open Books
    Our formal study of physics begins with kinematics which is defined as the study of motion without considering its causes.
  3. [3]
    [PDF] Physics Chapter Outline ch 1-3
    Equations of motion: - Assumes constant acceleration (neither magnitude or direction are changing) and straight line motion (acceleration is parallel to ...
  4. [4]
    [PDF] Physics, Chapter 2: Motion of a Particle (Kinematics)
    Hence Equations (2-11), (2-12), and. (2-13) are true in the general case where the motion is in any direction, not just parallel to one of the coordinate axes.<|control11|><|separator|>
  5. [5]
    [PDF] Numerical Solutions of Classical Equations of Motion - Physics
    Classical equations of motion, i.e., Newton's laws, govern the dynamics of systems ranging from very large, such as solar systems and galaxies, ...
  6. [6]
    [PDF] Classical Equations of Motion - Physics Courses
    Lagrange's equation of motion. Definition of momentum. Differential change in L. Legendre transform. Hamilton's equations of motion. Conservation of energy ...
  7. [7]
    The Feynman Lectures on Physics Vol. I Ch. 8: Motion
    If a body starts from rest and moves with a constant acceleration, g, its velocity v at any time t is given by v=gt.
  8. [8]
    ARISTOTLE, On the Heavens - Loeb Classical Library
    It is distinguished from the other four by its natural motion, which is circular, whereas theirs are rectilinear, either to or from the centre of the Universe.<|control11|><|separator|>
  9. [9]
    THE BOOK: ARISTOTLE'S PHYSICS
    According to Aristotle, the motion of physical bodies is of two types: natural motion and violent motion. Natural motion is the motion arising from the nature ...
  10. [10]
    Aristotle's Natural Philosophy
    May 26, 2006 · Nevertheless, when making this claim, Aristotle speaks about four kinds of motion and change only—those in substance, in quality, in ...
  11. [11]
    On the Heavens by Aristotle - The Internet Classics Archive
    On the Heavens by Aristotle, part of the Internet Classics Archive. ... heavier than wood. For all bodies, in spite of the general opinion to the contrary ...
  12. [12]
    Aristotle and Falling Objects | Diagonal Argument
    Aug 22, 2023 · Aristotle's physics is commonly said to state that heavier objects fall faster when every high-school kid should know they fall at the same speed.Observations · historical argument · Proportionality · Two kinds of air friction
  13. [13]
    [PDF] John Buridan and the Theory of Impetus - Fordham University Faculty
    Nov 23, 2006 · This impetus now [acting] together with its gravity moves it. Therefore, the motion becomes faster; and by the amount it is faster, so the ...
  14. [14]
    Medieval Theology II: Impetus | Being and Motion | Oxford Academic
    The Latin term impetus, regarding local motion, was introduced by Jean de Buridan (University of Paris, 1320–1358) in his Questions on the Eight Books of ...
  15. [15]
    The impetus theory: Between history of physics and science education
    Clagett, M. (ed.): 1968, Nicole Oresme and the Medieval Geometry of Qualities and Motions, University of Wisconsin Press, Madison. Google Scholar.
  16. [16]
    Nicole Oresme and Medieval Scientific Thought - jstor
    that Oresme uses the impetus theory to explain heavenly motion is the fact that he seems to as- sociate impetus with accelerated motion,23 while of course ...
  17. [17]
    Natural or violent motion? Galileo's conjectures on the fall of heavy ...
    According to Aristotelian physics, there was a fundamental distinction between natural and violent motion. When the cause of the motion was internal to the ...
  18. [18]
    Motion of Free Falling Object | Glenn Research Center - NASA
    Jul 3, 2025 · Galileo conducted experiments using a ball on an inclined plane to determine the relationship between the time and distance traveled.
  19. [19]
    Galileo's Experiments & Theory With Rolling Balls Down Inclined ...
    Galileo hypothesized that objects experienced uniform acceleration due to gravity. He devised an experiment involving balls rolling down an inclined plane to ...
  20. [20]
    Galileo's Acceleration Experiment
    Legend has it that Galileo performed this particular experiment from the leaning tower of Pisa.
  21. [21]
    Dialogues Concerning Two New Sciences | Online Library of Liberty
    To determine the momentum of a projectile at each particular point in its given parabolic path. Let bec be the semi-parabola whose amplitude is cd and whose ...
  22. [22]
  23. [23]
    Newton's Philosophiae Naturalis Principia Mathematica
    Dec 20, 2007 · The modern F=ma form of Newton's second law nowhere occurs in any edition of the Principia even though he had seen his second law formulated in ...Newton's Laws of Motion · Book 2 of the Principia · Book 3 of the Principia
  24. [24]
    Newton's Laws of Motion - Glenn Research Center - NASA
    Jun 27, 2024 · Sir Isaac Newton's laws of motion explain the relationship between a physical object and the forces acting upon it.Newton's First Law: Inertia · Newton's Second Law: Force · The Acceleration Of An...
  25. [25]
    Newton and Planetary Motion - UNL Astronomy
    In 1687 Isaac Newton published Philosophiae Naturalis Principia ... 2nd Law of Motion: F = ma. The acceleration of an object is proportional to the force acting ...
  26. [26]
    [PDF] A Historical Discussion of Angular Momentum and its Euler Equation
    As we will see in the following, Euler considered also the angular momentum of rigid bodies and proposed an equation for it in his works on mechanics. [10,11].
  27. [27]
    The history of the Méchanique analitique | Lettera Matematica
    Jun 6, 2014 · A description of the historical development of the ideas that led Lagrange to write his Méchanique analitique (1788).
  28. [28]
    Deriving Lagrange's equations using elementary calculus
    This paper provides a derivation of Lagrange's equations from the principle of least action using elementary calculus.
  29. [29]
    standard acceleration of gravity - CODATA Value
    standard acceleration of gravity $g_{\rm n}$ ; Numerical value, 9.806 65 m s ; Standard uncertainty, (exact).
  30. [30]
    1. Discovering Gravity - Galileo - The University of Virginia
    Aristotle was the first writer to attempt a quantitative description of falling motion: he wrote that an object fell at a constant speed, attained shortly after ...<|control11|><|separator|>
  31. [31]
    [PDF] 2-d Motion: Constant Acceleration - UF Physics Department
    Sep 6, 2012 · • Kinematic Equations of Motion (Vector Form). The velocity vector and position vector are a function of the time t. 2. 2. 1. 0. 0. )( tatvrtr о.
  32. [32]
    [PDF] Equation Number - KITP
    TABLE 1 VECTOR EQUATIONS FOR MOTION WITH CONSTANT ACCELERATION. Equation. Number. Equation. 11. V. 12. +at. +. 13a. #. V. 14. 15 rro + votat². Vo Vo+2a⚫ (r ro).
  33. [33]
    4.2: Acceleration Vector – University Physics Volume 1
    Suppose the acceleration function has the form a → ( t ) = ( a i ^ + b j ^ + c k ^ ) m/ s 2 , where a, b, and c are constants.
  34. [34]
    Projectile motion
    Reasoning: We have motion with constant acceleration in two dimensions, or projectile motion. The range of a projectile over level ground is R = (v02sin2θ0)/g.
  35. [35]
    3.4 Projectile Motion – College Physics - UCF Pressbooks
    The magnitudes of the components of the velocity v are v x = v cos θ and v y = v sin θ , where v is the magnitude of the velocity and θ is its direction, as ...
  36. [36]
    [PDF] CHAPTER 2 KINEMATICS OF A PARTICLE - Purdue Engineering
    The acceleration is now calculated, using the definition that it is the time-derivative of velocity: (Recall: these rates of change are. XYZ). • is directed ...
  37. [37]
    Motion With Constant Acceleration – Introductory Physics
    Every problem with constant acceleration can be solved using the kinematic equations. Conversely,. If an object does NOT have constant acceleration, you should ...
  38. [38]
    [PDF] Particle Kinematics and Space Curves 31 Mar 03 - UBC Math
    Mar 31, 2003 · Indeed, α = bT • u, β = bN • u, γ = bB • u. The Frenet-Serret Formulas. Recall that if u = u(t) obeys |u(t)| = 1 for all t, then u(t) ⊥ u. 0.
  39. [39]
    [PDF] 1. Phase space - UCLA Mathematics
    In classical mechanics, the phase space is the space of all possible states of a physical system; by “state” we do not simply mean the positions q of all ...
  40. [40]
    Differential Equations - Euler's Method - Pauls Online Math Notes
    Nov 16, 2022 · In this section we'll take a brief look at a fairly simple method for approximating solutions to differential equations.
  41. [41]
    Beyond velocity and acceleration: jerk, snap and higher derivatives
    Oct 13, 2016 · In this paper we will discuss the third and higher order derivatives of displacement with respect to time, using the trampolines and theme park roller coasters ...<|separator|>
  42. [42]
    5.3 Newton's Second Law - University Physics Volume 1 | OpenStax
    ### Summary of Newton's Second Law from OpenStax University Physics Volume 1, Section 5.3
  43. [43]
    15.1 Simple Harmonic Motion - University Physics Volume 1
    Sep 19, 2016 · We can use the equations of motion and Newton's second law ( F → net = m a → ) ( F → net = m a → ) to find equations for the angular frequency, ...
  44. [44]
    10.7 Newton's Second Law for Rotation - University Physics Volume 1
    Sep 19, 2016 · In this section, we introduce the rotational equivalent to Newton's second law of motion and apply it to rigid bodies with fixed-axis rotation.
  45. [45]
    Equations of motion for a rigid body - Engineering Mechanics
    Note: For a single particle the center of mass and the location of the particle are the same and one recovers Newton's second law.
  46. [46]
    [PDF] Chapter 4 Rigid Body Motion - Rutgers Physics
    In this chapter we develop the dynamics of a rigid body, one in which all interparticle distances are fixed by internal forces of constraint. This is,.Missing: source | Show results with:source
  47. [47]
    Moment of Inertia Tensor - Richard Fitzpatrick
    The moment of inertia tensor can be written as either a sum over separate mass elements, or as an integral over infinitesimal mass elements.
  48. [48]
    [PDF] 3D Rigid Body Dynamics: The Inertia Tensor - MIT OpenCourseWare
    The tensor of inertia gives us an idea about how the mass is distributed in a rigid body.
  49. [49]
    4.5: Euler's Equations of Motion - Physics LibreTexts
    Aug 7, 2022 · Euler's rotation equations are a vectorial quasilinear first-order ordinary differential equation describing the rotation of a rigid body, ...Missing: original | Show results with:original
  50. [50]
    [PDF] 8.01SC S22 Chapter 24: Physical Pendulum - MIT OpenCourseWare
    Mar 24, 2022 · A physical pendulum consists of a uniform rod of length d and mass m pivoted at one end. The pendulum is initially displaced to one side by ...
  51. [51]
    6.3: Lagrange Equations from d'Alembert's Principle
    Jun 28, 2021 · d'Alembert's Principle of virtual work. The Principle of Virtual Work provides a basis for a rigorous derivation of Lagrangian mechanics.Missing: 1750 | Show results with:1750
  52. [52]
    [PDF] Analytical Dynamics: Lagrange's Equation and its Application
    D'Alembert's principle may be stated by rewriting Equation (12) as miri − Fi = 0. (13). Taking the dot product of each of the Equations (13) with the ...
  53. [53]
    6.3: Motion Under the Action of a Central Force - Physics LibreTexts
    Apr 24, 2022 · Two important examples of central forces are (general) Newtonian gravity (2.2.2) and the Coulomb force (2.2.4) between two charged objects.
  54. [54]
    9.6: Conservation of Angular Momentum - Physics LibreTexts
    Nov 5, 2020 · Just as linear momentum is conserved when there is no net external forces, angular momentum is constant or conserved when the net torque is zero.
  55. [55]
    7.3 Conservation Laws and Symmetries
    That symmetry leads to the law of conservation of linear momentum. A system of particles in otherwise empty space conserves its total amount of linear momentum.Missing: Newtonian | Show results with:Newtonian
  56. [56]
    Mécanique analytique : Lagrange, J. L. (Joseph Louis), 1736-1813
    Jan 18, 2010 · Publication date: 1811 ; Topics: Mechanics, Analytic ; Publisher: Paris, Ve Courcier ; Collection: thomasfisher; universityofottawa; toronto; ...
  57. [57]
    [PDF] The Lagrangian Method
    The Euler-Lagrange equations are valid in any coordinates. Note that the above proof did not in any way use the precise form of the Lagrangian. If. L were equal ...
  58. [58]
    [PDF] Energy Methods: Lagrange's Equations - MIT OpenCourseWare
    A significant advantage of the Lagrangian approach to developing equations of motion for complex systems comes as we leave the cartesian xi coordinate system ...
  59. [59]
  60. [60]
    [PDF] Double pendulum: An experiment in chaos - RB Levien
    The double pendulum experiment demonstrates chaos through sensitive dependence on initial conditions, and can also show small angle and zero gravity motion.
  61. [61]
    English trans. of E. Noether Paper - UCLA
    Wiss. zu Göttingen 1918, pp235-257. English translation: M.A. Tavel, Reprinted from "Transport Theory and Statistical Mechanics" 1(3), 183 ...
  62. [62]
    [PDF] ON A GENERAL METHOD IN DYNAMICS By William Rowan Hamilton
    This edition is based on the original publication in the Philosophical Transactions of the. Royal Society, part II for 1834. The following errors in the ...
  63. [63]
    Hamiltonian systems - Scholarpedia
    Aug 19, 2007 · In 1834 William Rowan Hamilton showed that Newton's equations F = ma for a set of particles in a conservative force field F = -\nabla V with " ...
  64. [64]
    Versuch einer Theorie der electrischen und optischen ...
    Jan 18, 2008 · Versuch einer Theorie der electrischen und optischen Erscheinungen in ... by: Hendrik Antoon Lorentz ... PDF download · download 1 file · SINGLE ...
  65. [65]
    Charged Particle in a Magnetic Field - Richard Fitzpatrick
    A charged particle placed in a magnetic field executes a circular orbit in the plane perpendicular to the direction of the field.
  66. [66]
    [PDF] 5. Electromagnetism and Relativity - DAMTP
    The relativistic formulation of the Lorentz force (5.30) also contains an extra equation coming from µ = 0. This reads. dP0 d⌧. = q. cE · u. (5.33). Recall that ...
  67. [67]
    Abraham–Lorentz versus Landau–Lifshitz - AIP Publishing
    Apr 1, 2010 · The classical Abraham–Lorentz formula for the radiation reaction on a point charge suffers from two notorious defects: runaways and preacceleration.
  68. [68]
    A Universal Model: The Vlasov Equation - Taylor & Francis Online
    See also the classical textbook by Chandrasekhar Citation(1942), for a historical review on the Vlasov equation the reader is also refered to a paper by M.
  69. [69]
    [PDF] Synchrotron Radiation - SLAC National Accelerator Laboratory
    Dec 2, 2001 · Particle Beam Optics equation of motion therefore defines the reference orbit for particles with en- ergy E =Eo(1 +δ). Such particles ...
  70. [70]
    [PDF] Space and Time - UCSD Math
    Minkowski had not understood Einstein's 1905 paper on special relativity. I ... is an absolute four-dimensional world is contained in Minkowski's paper it-.
  71. [71]
    [PDF] ON THE ELECTRODYNAMICS OF MOVING BODIES - Fourmilab
    It is essential to have time defined by means of stationary clocks in the stationary system, and the time now defined being appropriate to the stationary system ...Missing: proper | Show results with:proper
  72. [72]
    [PDF] The theory of the rigid electron in the kinematics of the principle of ...
    Born – The theory of the rigid electron in the kinematics of the relativity ... accelerated motion, and every motion can be approximated by hyperbolic motions.
  73. [73]
    [PDF] Notes
    THE PRINCIPLE OF CONSERVATION. OF MOTION. OF THE CENTER. OF GRAVITY. AND. THE ... of the center of gravity to be valid (at least in first approximation) also.
  74. [74]
    [PDF] Short Overview of Special Relativity and Invariant Formulation of ...
    Abstract. The basic concepts of special relativity are presented in this paper. Con- sequences for the design and operation of particle accelerators are ...
  75. [75]
    Spinning test-particles in general relativity. I - Journals
    The equations of motion of spinning test particles are derived. The transformation properties are discussed and the equations of motion are written in a ...
  76. [76]
    Dynamics of extended bodies in general relativity. I. Momentum and ...
    The paper proposes definitions for total momentum and spin tensor of an extended body, and defines total rest energy as the length of the momentum vector.
  77. [77]
    Momentum-velocity relation for the Mathisson-Pirani spin condition
    Apr 16, 2018 · The Mathisson-Papapetrou-Dixon (MPD) equations, providing the ... The spin evolution equation becomes the Fermi-Walker transport law (e.g.
  78. [78]
    Spinning test particles in a Kerr field - Oxford Academic
    Mathisson—Papapetrou equations are solved numerically to obtain trajectories of spinning test particles in (the meridional section of) the Kerr space—time.
  79. [79]
    Resonant orbits for a spinning particle in Kerr spacetime | Phys. Rev. D
    Jun 22, 2020 · We start with the Mathisson-Papapetrou equations under the linear spin approximation and primarily concentrate on two particular events: first, ...
  80. [80]
    Black holes merge with strongly inclined rotation axes
    Oct 28, 2025 · Rotating black holes deform: Thanks to a scientist's analysis of gravitational waves, it has been confirmed for the first time that black holes ...
  81. [81]
    Classical Wave Equations - Galileo
    We derive the wave equation from F=ma for a little bit of string or sheet. The equation corresponds exactly to the Schrödinger equation for a free particle with ...Introduction · Simple Solutions to the... · The Schrödinger Equation for...
  82. [82]
    [PDF] DERIVATION AND ANALYSIS OF SOME WAVE EQUATIONS
    In most cases, one can start from basic physical principles and from these derive partial differential equations (PDEs) that govern the waves. In Section 4.2 ...
  83. [83]
    47 Sound. The wave equation - Feynman Lectures - Caltech
    This behavior of electric fields may be described by saying that if f1(x−ct) is a wave, and if f2(x−ct) is another wave, then their sum is also a wave. This is ...
  84. [84]
    Sound Waves - Galileo
    From F = ma to the Wave Equation​​ Having found how the local pressure variation relates to s(x,t), we're ready to derive the wave equation from F=ma for a thin ...
  85. [85]
    [PDF] Maxwell's Equations and EM Waves - UF Physics Department
    Nov 21, 2006 · Derivation of Electromagnetic Wave Equation. Now let's see how we can combine the differential forms of Maxwell's equations to derive a set ...
  86. [86]
    Dispersion relation - MIT
    The dispersion relation ω ( K ) \omega(\mathcal K) ω(K) relates angular frequency with wavenumber. In a linear dispersion relation, ω = v p K \omega = v_p ...
  87. [87]
    [PDF] 11. Status of Higgs Boson Physics - Particle Data Group
    May 31, 2024 · Eleven years after the discovery, substantial progress in the field of Higgs boson physics has been accomplished and a significant number of ...
  88. [88]
    Unveiling new phases of the Standard Model Higgs potential
    Jan 31, 2025 · We present evidence for new phases of the Standard Model Higgs potential. We study the Standard Model physical trajectory accounting for the Higgs curvature ...
  89. [89]
    [PDF] 1926-Schrodinger.pdf
    It was stated in the beginning of this paper that in the present theory both the laws of motion and the quantum conditions can be deduced from one Hamiltonian ...