Fact-checked by Grok 2 weeks ago

Test particle

In physics, a test particle is an idealized model of an object whose , charge, or other physical are assumed to be negligible, except insofar as they allow the object to interact with an external ; this enables the particle to probe the 's effects without significantly altering the field itself. The concept is fundamental in theories, where the test particle's reveals of the surrounding gravitational or electromagnetic environment. In and , the test particle is often referred to as a test charge, typically taken as a small positive charge q_0 that experiences a force \mathbf{F} = q_0 \mathbf{E} in an \mathbf{E}, allowing the field to be defined operationally as \mathbf{E} = \mathbf{F}/q_0 without the test charge contributing to the field's source. This idealization assumes the test charge's own field is insignificant compared to the external one, ensuring the measurement reflects the undisturbed field configuration. The direction of the is conventionally defined as that of the force on a positive test charge, pointing away from positive sources and toward negative ones. In gravitational contexts, a test particle is an object with sufficiently small mass that its gravitational influence on surrounding bodies is negligible, permitting it to follow geodesics or orbits that trace the ambient without back-reaction. For instance, in Newtonian gravity, test particles of varying composition fall with the same in the same location, a experimentally verified from Galileo's Pisa demonstrations to modern space-based tests, such as the mission (2016–2018), showing equivalence within 1 part in $10^{15}. This weak underpins the use of test particles in , where they define local inertial frames by freely falling along geodesics. The test particle approximation extends to plasma physics, relativistic dynamics, and numerical simulations, where it simplifies calculations by treating the particle as passive in complex field environments, such as magnetized plasmas or black hole spacetimes. While idealized, the approximation's validity depends on the scale: for example, a 10 kg mass may not qualify as a test particle near lighter objects if it induces measurable displacements within experimental tolerances.

Fundamentals

Definition

In physics, a test particle is an idealized model of an object whose physical properties—such as , charge, , or extent—are assumed to be negligible except for the specific attribute relevant to the external field under study, ensuring it does not significantly influence or back-react on that field. This allows the test particle to serve as a passive probe, tracing the behavior and structure of fields like electric or gravitational ones without altering their configuration. By treating the particle as point-like and non-self-interacting, the model simplifies the analysis of field-induced motion, focusing solely on the external influences. Unlike real particles, which possess finite properties that can mutually interact and modify surrounding fields, a test particle is hypothetical and designed to isolate the effects of the ambient field on its trajectory. This distinction is crucial in theoretical frameworks, where the test particle's response—such as acceleration or deflection—reveals field characteristics without complications from self-gravity, self-charge repulsion, or other feedback effects. Representative examples illustrate its utility: in , a test charge is a small positive charge placed in an to measure the force per unit charge, thereby mapping the field's direction and magnitude without disturbing the source charges. Similarly, in Newtonian , a test mass is a body with such minimal mass that it experiences from external sources while causing negligible attraction on them, enabling the definition and exploration of the . These applications underscore the test particle's role in deriving field properties from observed forces on the idealized object.

Assumptions and Limitations

The test particle approximation in physics relies on several key assumptions to simplify the analysis of motion in external fields. Primarily, the test particle is assumed to have negligible (approaching m \to 0) or charge, ensuring it does not produce significant , back-reaction on the source, or alteration of the external field through its own field contributions. In , this means the test charge q is taken to be infinitesimally small so that it does not disturb the charge distribution creating the , allowing the field to be defined as the force per unit charge \mathbf{E} = \lim_{q \to 0} \frac{\mathbf{F}}{q}. Similarly, in Newtonian , the test mass is assumed negligible to avoid perturbing the generated by source masses, with the field defined as \mathbf{g}(\mathbf{r}) = \lim_{m \to 0} \frac{\mathbf{F}}{m}. In relativistic contexts, such as , the test particle follows a in the without contributing to , again under the negligible limit. Another fundamental assumption is the point-particle limit, where the test particle is idealized as having zero spatial extent. This simplifies trajectory calculations by neglecting effects from the particle's internal structure, such as tidal deformations or multipole moments, which could otherwise complicate the motion in non-uniform fields. Additionally, for charged test particles, negligible couplings beyond the primary interaction (e.g., minimal magnetic effects in electrostatic approximations) are assumed to isolate the dominant . These assumptions hold best in regimes of weak external fields, low velocities relative to the , and classical scales where quantum effects are irrelevant. Despite their utility, these assumptions impose significant limitations on the test particle model's validity. The approximation breaks down when the test particle's own field becomes comparable to the external one, such as in close proximity to or in scenarios with high particle density, leading to non-negligible back-reaction or feedback effects like return currents in environments. For instance, in gravitational two-body problems, treating one body as a test particle fails if the masses are comparable, requiring full N-body dynamics instead. In strong-field regimes, such as near black holes or in high-energy collisions, quantum corrections (e.g., from or ) render the classical point-particle description inadequate, necessitating more advanced treatments like on curved spacetimes. The point-like assumption also overlooks extended-body effects, like forces, which become important for real particles with finite size, limiting applicability to idealized scenarios. Overall, the model is most reliable in dilute, weak-field limits but must be supplemented or abandoned in dense, strong, or quantum-dominated systems.

Historical Development

Origins in Classical Physics

The concept of a test particle emerged implicitly in through approximations in gravitational and electrostatic interactions, where one body's influence on the field generated by another is neglected due to its negligible or charge. In Newtonian gravity, treated orbiting bodies, such as planets around the Sun, as moving in a central produced by a much more massive central body, disregarding the mutual perturbations among the lighter objects to simplify the analysis of orbital motion. This approach, outlined in Newton's (1687), laid the groundwork for viewing smaller masses as probes that trace the without altering it. In , the idea took shape through Charles-Augustin de Coulomb's experiments in 1785, where he measured the repulsive between small, lightweight charged balls using a torsion balance, assuming the objects' minimal size and mass prevented significant mutual disturbance of their charge distributions. These balls effectively served as test charges, allowing to establish the of electric without the complications of back-reaction from the probe itself. The test particle notion gained formal structure in 19th-century field theories, particularly through Michael Faraday's conceptualization of electric and magnetic fields as continuous media defined by the forces they exert on hypothetical small charges or magnets, and James Clerk Maxwell's mathematical unification of these ideas in his treatise on . Faraday's qualitative lines of force and explicitly incorporated test charges to define field strengths, shifting from action-at-a-distance to field-mediated interactions where the probe's effect on the source is idealized as zero. By the early , the test particle approximation was refined as a standard pedagogical and analytical tool in for solving problems in external fields, such as central force motions, enabling the inference of field properties from observed particle trajectories in inverse problems. This usage persisted in texts emphasizing approximations for systems with disparate masses or charges, bridging classical formulations toward broader theoretical extensions.

Adoption in Relativistic Theories

The concept of the test particle emerged implicitly in through the treatment of charged particles under Lorentz transformations, where their motion in electromagnetic fields was analyzed without significant backreaction on the fields themselves. In Einstein's paper on the electrodynamics of moving bodies, the trajectories of such particles were described via the invariance of , laying groundwork for relativistic particle dynamics, though the explicit distinction of test particles as negligible-mass probes was not yet formalized. This implicit adoption facilitated early calculations of relativistic orbits and forces, bridging to relativistic kinematics. The key development occurred with the advent of in , where Einstein's positioned freely falling test particles—treated as having negligible mass and thus no gravitational backreaction—as following in curved . This principle equated inertial and gravitational mass, allowing test particles to serve as ideal probes of geometry without altering the metric. Einstein's formulation in his 1916 review article explicitly invoked this for deriving the in gravitational fields, marking the transition from classical point masses to relativistic test bodies. Concurrently, Hilbert's action principle for incorporated matter terms that supported geodesic motion for test particles, providing a variational framework for their paths in curved backgrounds. Formalization advanced in the and through mathematical refinements, notably Levi-Civita's development of affine connections and , which rigorously defined for test particles in arbitrary curved . His 1917 work on the geometric specification of Riemannian curvature enabled precise descriptions of how test particles trace spacetime structure without self-interaction. This era saw the concept solidify as essential for solving the independently of the full field equations. A comprehensive modern review by in 2004 traces this evolution, emphasizing point particles as scalar, electromagnetic, or gravitational charges moving in prescribed backgrounds. Significant milestones included extensions to more complex cases, such as Mathisson's 1937 equations and Papapetrou's 1951 formulation, which adapted the test particle framework to include spin effects while preserving the negligible-mass approximation. These developments, building on earlier variational principles, influenced subsequent applications by decoupling particle trajectories from field sourcing. The adoption of test particles thus enabled pivotal advances in physics—via solutions like Schwarzschild's 1916 metric, probed by infalling test bodies—and cosmological models, where they trace large-scale without requiring complete nonlinear solutions.

Applications in Classical Physics

Electrostatics

In electrostatics, the \mathbf{E} at a point in space is operationally defined as the electrostatic force \mathbf{F} experienced by a small positive charge q placed at that point, divided by the magnitude of the charge: \mathbf{E} = \frac{\mathbf{F}}{q}. The charge must be infinitesimally small to avoid perturbing the existing field significantly, ensuring the measurement reflects the external field alone./Volume_B:_Electricity_Magnetism_and_Optics/B03:_The_Electric_Field_Due_to_one_or_more_Point_Charges) This definition allows the to be treated as a vector quantity independent of the charge, with direction indicating the force on a positive charge and magnitude in newtons per coulomb (N/C). For a point source charge Q in vacuum, the electric field at a distance r from the charge is radial and given by \mathbf{E} = \frac{1}{4\pi\epsilon_0} \frac{Q}{r^2} \hat{r}, where \epsilon_0 is the vacuum permittivity, \hat{r} is the unit vector in the radial direction, and the field points away from Q if positive or toward it if negative. This expression derives from Coulomb's law and embodies the inverse-square nature of electrostatic interactions./Volume_B:_Electricity_Magnetism_and_Optics/B03:_The_Electric_Field_Due_to_one_or_more_Point_Charges) Under the of a , a charge initially at follows a straight-line to \mathbf{E}, accelerating constantly with a = \frac{qE}{m}, where m is the charge's mass./06:_General_Planar_Motion/6.03:_Motion_Under_the_Action_of_a_Central_Force) In the non-uniform field of a point charge, assuming negligible magnetic effects and initial velocity, the trajectory is a conic section: hyperbolic for like-charge repulsion or elliptic for opposite-charge attraction if bound./06:_General_Planar_Motion/6.03:_Motion_Under_the_Action_of_a_Central_Force) Test charges provide the observational basis for , which relates the through a closed surface to the enclosed charge: \oint \mathbf{E} \cdot d\mathbf{A} = \frac{Q_{\text{enc}}}{\epsilon_0}. By placing test charges at various points and noting field symmetry around symmetric charge distributions, such as spheres, the law emerges from the consistent inverse-square pattern. In practice, high-impedance voltmeters serve as real-world analogs to test charges, drawing negligible current to measure potential without significantly altering the electrostatic field. The vacuum permittivity \epsilon_0 has an SI value of $8.854 \times 10^{-12} F/m, scaling the strength of electrostatic forces in the meter-kilogram-second-ampere system. In contrast, the cgs electrostatic unit system absorbs \frac{1}{4\pi\epsilon_0} into the charge definition, using statcoulombs and dynes for a more compact but dimensionally distinct formulation.

Newtonian Gravity

In Newtonian gravity, a test particle is defined as an object with sufficiently small that its gravitational influence on surrounding bodies is negligible, allowing its motion to be determined solely by the external without back-reaction effects. The force on such a test mass m in a gravitational field \mathbf{g} is given by \mathbf{F} = m \mathbf{g}, where for a point mass M at the origin, the field is \mathbf{g} = -\frac{[G](/page/G) M}{r^2} \hat{\mathbf{r}}, with G the and \hat{\mathbf{r}} the unit vector in the radial direction. This formulation arises directly from , which states that every particle attracts every other with a force proportional to the product of their masses and inversely proportional to the square of the distance between them. The full two-body gravitational interaction between masses m_1 and m_2 separated by \mathbf{r}_1 - \mathbf{r}_2 is \mathbf{F} = -\frac{[G](/page/G) m_1 m_2}{|\mathbf{r}_1 - \mathbf{r}_2|^2} \hat{\mathbf{r}}, where \hat{\mathbf{r}} points from m_1 to m_2..pdf) In the test particle approximation, one sets m_2 \to 0 while keeping m_1 = M fixed, reducing the problem to the motion of a single body in the field of the dominant and simplifying the to a one-body central problem..pdf) This reduction is achieved by transforming to the center-of-mass frame, where the relative motion equation becomes \frac{d^2 \mathbf{r}}{dt^2} = -\frac{[G](/page/G) M}{r^3} \mathbf{r}, with M = m_1 + m_2 \approx m_1. For central forces like gravity, the orbital motion can be analyzed using an effective potential V_{\text{eff}}(r) = -\frac{G M m}{r} + \frac{L^2}{2 m r^2}, where L is the conserved angular momentum, combining the gravitational potential with the centrifugal term to govern radial motion as an effective one-dimensional problem. Solutions to this system yield Keplerian orbits: bound elliptical paths for negative total energy, parabolic trajectories for zero energy, and hyperbolic escapes for positive energy, all conic sections with the central mass at one focus..pdf) A practical example is the motion of Earth-orbiting satellites, where the satellite's mass (typically kilograms) is negligible compared to Earth's ($5.97 \times 10^{24} kg), allowing accurate prediction via Keplerian elements without accounting for mutual perturbations. This holds well when masses differ significantly but fails for comparable masses, such as systems, where the \mu = \frac{m_1 m_2}{m_1 + m_2} must be used to describe the equivalent one-body problem orbiting the total mass..pdf) The test particle limit in gravity shares a structural with the test charge in , both relying on inverse-square laws for field-induced motion.

Applications in General Relativity

Geodesic Motion

In , the trajectory of a test particle under the influence of alone is described as geodesic motion, where the particle follows the "straightest" possible path in curved , free from non-gravitational forces. This path, known as a , generalizes the concept of a straight line from to the of , with deviations from flat-space motion quantified by the derived from the . The equation of motion for such a test particle is the geodesic 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 x^\mu are the spacetime coordinates, \tau is the proper time parameterizing the worldline for timelike geodesics, and \Gamma^\mu_{\alpha\beta} are the Christoffel symbols of the second kind encoding the spacetime curvature. This equation arises from the requirement that the tangent vector to the geodesic is parallel-transported along the curve, ensuring no acceleration relative to local inertial frames. The spacetime curvature itself is governed by the Einstein field equations, G_{\mu\nu} = \frac{8\pi G}{c^4} T_{\mu\nu}, which connect the Einstein tensor G_{\mu\nu}—a measure of —to the stress-energy tensor T_{\mu\nu} representing mass-energy distributions; for a test particle approximation, the particle's negligible means its own T_{\mu\nu} is ignored, so it probes the background geometry created by other sources. Specific examples illustrate motion in realistic scenarios. In the , which describes the exterior to a non-rotating, spherically symmetric such as a black hole or star, timelike geodesics include radial infall paths where particles accelerate toward the central singularity and bound orbits exhibiting precession, as seen in the anomalous advance of Mercury's perihelion by 43 arcseconds per century. Null geodesics, corresponding to massless particles like photons with affine parameter replacing proper time, manifest as the bending of light around massive bodies, deflecting starlight by 1.75 arcseconds when grazing the Sun's limb. Geodesic motion embodies the , positing that the effects of gravity are indistinguishable from acceleration in locally inertial frames, thereby interpreting gravitation not as a force but as the inertial response to . This framework unifies gravitational phenomena across scales, from planetary orbits to cosmological expansion, and underpins predictions verified by observations like gravitational lensing and GPS corrections.

Effects of Spin

In the standard test particle approximation within , the particle is assumed to have negligible , treating it as a point-like object with zero intrinsic , such that its worldline follows a path determined solely by the . This scalar or is considered infinitesimal compared to the particle's and the gravitational field's , simplifying the to the equation without additional forces. To incorporate spin effects, the model is extended to describe a spinning test particle using the Mathisson-Papapetrou-Dixon (MPD) equations, which account for the coupling between the particle's spin tensor and the . These equations govern the evolution of the particle's linear p^\mu and antisymmetric spin tensor S^{\mu\nu}: \frac{D p^\mu}{d\tau} = -\frac{1}{2} R^\mu_{\ \nu\rho\sigma} u^\nu S^{\rho\sigma}, \frac{D S^{\mu\nu}}{d\tau} = p^\mu u^\nu - p^\nu u^\mu, where \tau is the proper time, u^\mu is the four-velocity, R^\mu_{\ \nu\rho\sigma} is the Riemann curvature tensor, and \frac{D}{d\tau} denotes covariant differentiation along the worldline. A supplementary condition, such as the Tulczyjew condition S^{\mu\nu} p_\nu = 0, is imposed to fix the worldline and ensure physical consistency. In the limit of vanishing spin, S^{\mu\nu} \to 0, the equations reduce to geodesic motion with p^\mu = m u^\mu for rest mass m. The introduces two primary effects: deviation from motion and . The right-hand side of the momentum equation represents the Mathisson force, a non- proportional to the tensor and , causing the worldline to curve away from the by an amount scaling as S / (m R), where R is a radius. For the evolution, the MPD equations imply that S^{\mu\nu} undergoes Fermi-Walker along the worldline, meaning the vector precesses due to the particle's and the , without torque in the particle's . This includes de Sitter (geodetic) and Lense-Thirring () components, altering the particle's orientation relative to distant observers. These effects are applied in astrophysical contexts involving compact objects. In neutron star environments, spinning test particles model pulsar emissions or debris orbits, where spin-curvature coupling influences stability and rates around the star's magnetic and gravitational fields. For accretion disks, the MPD framework describes how spinning particles deviate from Keplerian orbits, affecting disk and transport near the event horizon. In extreme mass-ratio inspirals (EMRIs), such as a stellar-mass or spiraling into a , spin modulates waveforms, enabling detection of secondary spin parameters with future observatories like . The equations are valid under the pole-dipole approximation, assuming small magnitudes (|S| \ll m^2 in natural units) and classical regimes where quantum effects are negligible, as they do not incorporate the for fermionic . They break down for large or Planck-scale curvatures, requiring higher-multipole extensions or quantum treatments. Historically, the equations emerged from foundational work in the mid-20th century: Mathisson introduced early forms in 1937, Papapetrou formalized the spinning particle dynamics in 1951, Tulczyjew refined the supplementary conditions in 1959, and Dixon extended the framework to multipolar bodies in the 1960s–1970s.

References

  1. [1]
    test particles - Einstein-Online
    test particles In the context of gravity: body whose mass is so small, that it can be used to probe the gravitational influences of their bodies.
  2. [2]
  3. [3]
    [PDF] Chapter 15.1-2 The Electric Field - General Physics II
    property only of and not of other charges we place in its vicinity. • We define the electric field in terms of the electric force that acts on a “test charge”:.
  4. [4]
    Electric field - HyperPhysics Concepts
    The direction of the field is taken to be the direction of the force it would exert on a positive test charge. The electric field is radially outward from a ...
  5. [5]
    [PDF] 18.6 The Electric Field - 4.1 The Concepts of Force and Mass
    NOTE: It is the surrounding charges that create the electric field at a given point. The effect of the test charge itself is NEVER included in this definition. ...
  6. [6]
    18.3 Electric Field - Physics | OpenStax
    Mar 26, 2020 · Now consider placing a test charge in the field. A test charge is a positive electric charge whose charge is so small that it does not ...
  7. [7]
    On the unreasonable effectiveness of the post-Newtonian ... - PMC
    A “test particle”—that is, a particle whose mass is sufficiently small that its own contribution to the curvature of spacetime can be ignored—moves on a ...
  8. [8]
    [PDF] Newtonian Gravity - Assets - Cambridge University Press
    To recover the force, just multiply the field by m: F = mg. The mass m at r will play the role of a “test mass” just as we often think of a positive “test ...
  9. [9]
    Test-particle dynamics in general spherically symmetric black hole ...
    May 18, 2018 · In this paper, a mathematical framework to describe test-particle dynamics in general black-hole spacetimes is presented and subsequently used ...<|control11|><|separator|>
  10. [10]
    Test particle sampling and particle acceleration in a 2D coronal ...
    Without background collisions, the test particle simulations rely on two strong assumptions: collisionless motion and negligible particle feedback. Report issue ...
  11. [11]
    Gravity - Newton's Law, Universal Force, Mass Attraction | Britannica
    Sep 29, 2025 · Newton assumed the existence of an attractive force between all massive bodies, one that does not require bodily contact and that acts at a distance.
  12. [12]
    Newton's Philosophiae Naturalis Principia Mathematica
    Dec 20, 2007 · The rules for calculating orbital motion that Kepler put forward in the first two decades of the seventeenth century had indeed achieved a ...
  13. [13]
    June 1785: Coulomb Measures the Electric Force
    Jun 1, 2016 · Coulomb became interested in measuring the electrical force between small charged objects and perfected a torsion balance which could reliably ...<|control11|><|separator|>
  14. [14]
    Coulomb force | Electric Charge, Interaction & Physics - Britannica
    Oct 31, 2025 · Coulomb force, attraction or repulsion of particles or objects because of their electric charge. One of the basic physical forces.
  15. [15]
    The conceptual origins of Maxwell's equations and gauge theory
    Nov 1, 2014 · Already in Faraday's electrotonic state and Maxwell's vector potential, gauge freedom was an unavoidable presence.
  16. [16]
    Electromagnetism - Induction, Faraday, Magnetism | Britannica
    Sep 26, 2025 · Maxwell's four field equations represent the pinnacle of classical electromagnetic theory. Subsequent developments in the theory have been ...
  17. [17]
    Newtonian Gravity with a Tiny Test Mass | Science 2.0
    Jan 11, 2014 · The tiny test mass is so small it does not materially change the field equations above. The action is a function of a parameterized position and ...<|separator|>
  18. [18]
    Electric field (article) | Electrostatics - Khan Academy
    The electric field is normalized electric force. Electric field is the force experienced by a test charge that has a value of + 1.
  19. [19]
    6.2 Explaining Gauss's Law - University Physics Volume 2 | OpenStax
    Oct 6, 2016 · Figure 6.16 The electric flux through any closed surface surrounding a point charge q is given by Gauss's law. (a) Enclosed charge is positive.
  20. [20]
    5 Application of Gauss' Law - The Feynman Lectures on Physics
    According to Gauss' law, the total flux of E from this surface is equal to the charge inside divided by ϵ0. Since the field is assumed to be normal to the ...5--2equilibrium In An... · 5--6a Sheet Of Charge; Two... · 5--8is The Field Of A Point...
  21. [21]
    [PDF] A New Class of Non Loading Voltmeters for Contacting and ...
    By necessity, all Voltmeters have high input impedance, but in some applications which require minimum loading, exceptionally high input impedance is required.
  22. [22]
    vacuum electric permittivity - CODATA Value
    vacuum electric permittivity $\varepsilon_0$ ; Numerical value, 8.854 187 8188 x 10-12 F m ; Standard uncertainty, 0.000 000 0014 x 10-12 F m ...Missing: SI | Show results with:SI
  23. [23]
    3. Electric and Magnetic Units: Gaussian and SI - Galileo and Einstein
    Gaussian units use cgs (cm, g, s) and dyne, while SI uses current (I) and has different dimensions for electric and magnetic fields. Gaussian is more ...
  24. [24]
    Newton's Law of Gravitation Derivation
    Dec 11, 2018 · The value of "G" was first measured by Henry Cavendish in 1798. The currently accepted value of "G" is 6.672 x 10-11 N2/kg2. Newton's model ...
  25. [25]
    [PDF] Keplerian Orbits Isaac Newton derived his now classical differential ...
    Jun 21, 2016 · The path traced by z(τ) is a lighter particle's orbit around a heavier attracting body and lies in a plane that includes their center of mass, ...
  26. [26]
    [PDF] 4. Central Forces - DAMTP
    In this section we will study the three-dimensional motion of a particle in a central force potential. Such a system obeys the equation of motion.
  27. [27]
    Spinning test-particles in general relativity. I - Journals
    A method for the derivation of the equations of motion of test particles in a given gravitational field is developed.
  28. [28]
    Spinning test particles in the spacetime | Phys. Rev. D
    Nov 26, 2019 · We derive the equations of motion for spinning test particles by using the Mathisson-Papapetrou-Dixon equations together with the Tulczyjew spin ...
  29. [29]
    On the motion of spinning test particles in plane gravitational waves
    Aug 14, 2003 · The Mathisson-Papapetrou-Dixon equations for a massive spinning test particle in plane gravitational waves are analysed and explicit solutions ...
  30. [30]
    [1110.1967] Mathisson-Papapetrou-Dixon equations in the ... - arXiv
    Oct 10, 2011 · The Mathisson-Papapetrou-Dixon equations describe the motion of a spinning test particle in Schwarzschild and Kerr backgrounds, with a new ...Missing: 1960s | Show results with:1960s
  31. [31]
    Fermi–Walker transport and Thomas precession - IOPscience
    An exact derivation of the Thomas precession formula is presented based on the Fermi–Walker transport equation.Missing: GR | Show results with:GR
  32. [32]
    [gr-qc/0202085] Circular Holonomy, Clock Effects and ... - arXiv
    Feb 24, 2002 · The historical origins of Fermi-Walker transport and Fermi coordinates and the construction of Fermi-Walker transported frames in black hole ...
  33. [33]
    Spin effects on neutron star fundamental-mode dynamical tides
    Aug 9, 2021 · ... relativity simulations of binary neutron star and neutron star–black-hole binaries. ... Hawking, Black holes in general relativity, Commun. Math.
  34. [34]
    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.
  35. [35]
    Generalized Mathisson-Papapetrou-Tulczyjew-Dixon Equations
    Apr 25, 2022 · We derive two generalizations of Mathisson-Papapetrou-Tulczyjew-Dixon equations from Casalbuoni-Brink-Schwarz type pseudoclassical Lagrangians of Majorana ...
  36. [36]
    Equations of the Dynamics of Spinning Particles in General Relativity
    Feb 28, 2020 · We analyze the role of the Mathisson–Papapetrou equations in establishing the regularities of spingravitational interaction under the ...
  37. [37]
    Republication of: New mechanics of material systems - ResearchGate
    In this form the equations are often referred to as the "Mathisson-Papapetrou-Dixon" equations (MPEQs), emphasizing the important early contribution of ...
  38. [38]
    Dynamics of extended bodies in general relativity. I. Momentum and ...
    Definitions are proposed for the total momentum vector pα and spin tensor Sαβ of an extended body in arbitrary gravitational and electromagnetic fields.