Fact-checked by Grok 2 weeks ago

Potential energy

Potential energy is the energy possessed by an object or due to its position within a force field or the arrangement of its internal components, representing stored that can be converted into other forms such as . This form of energy arises from conservative forces, where the work done by the force depends only on the initial and final positions, not the path taken, allowing for a well-defined potential energy . The unit of potential energy is the joule (J), equivalent to a newton-meter, reflecting its role as the work required to assemble or position the against the conservative force. Key types of potential energy include gravitational potential energy, which depends on an object's and height relative to a reference point in Earth's gravitational field, calculated as U_g = mgh where m is , g is the (approximately 9.8 m/s²), and h is height. For instance, a 75-kg object elevated 10 meters above the ground stores about 7,350 J of gravitational potential energy, which can be released as the object falls. Another prominent form is elastic potential energy, stored in stretched or compressed elastic materials like springs, given by U_s = \frac{1}{2} k x^2 where k is the spring constant and x is the displacement from . This energy powers devices such as catapults or shock absorbers, converting back to upon release. Beyond mechanical forms, potential energy encompasses electric, magnetic, chemical, and varieties, each tied to specific interactions like positions in or molecular bond configurations. The force associated with potential energy is the negative gradient of the , \mathbf{F} = -\nabla U, ensuring that decreases in potential energy correspond to increases in , as governed by the conservation of in isolated systems. This principle underpins applications in physics, , and everyday phenomena, from hydroelectric power generation to biochemical reactions.

Overview and History

Definition and Basic Concepts

Potential energy, denoted as U, is the energy that a physical system possesses due to its position within a force field or its internal of particles, that can be converted into other forms of , such as , without dissipation when only conservative forces are present. This stored represents the capacity of the system to perform work when its changes, arising fundamentally from interactions governed by conservative forces, where the work done by the force is independent of the path taken. In essence, potential energy quantifies the potential for motion or change inherent in the system's arrangement relative to other objects or fields. A key basic concept is that potential energy emerges specifically from conservative forces, such as or electrostatic forces, which allow the definition of a because the net work done in moving between two points depends only on the endpoints, not the . For intuitive understanding, consider a held above the ground: its potential energy stems from its elevated position in Earth's , ready to convert to if released. Similarly, a compressed or stretched stores potential energy due to the elastic deformation in its molecular configuration, which can propel an attached object upon release. These examples illustrate how potential energy is "stored" temporarily, contrasting with , which is associated with the system's motion. In isolated systems subject only to conservative forces, the total —the sum of potential energy U and K—remains , as interconversions between U and K occur without loss to other forms like . This principle, U + K = \text{[constant](/page/Constant)}, underscores the interchangeable nature of these energy components in such systems. The specific mathematical expression for potential energy in any given scenario depends on the underlying force law, such as those for or elasticity, which will be explored in subsequent sections.

Historical Background

The concept of potential energy has deep roots in , tracing back to Aristotle's notion of energeia, which described the actualization of potentiality in natural processes and motion, providing an early framework for understanding stored capacities in physical systems. This idea evolved significantly in the with Gottfried Wilhelm Leibniz's introduction of vis viva in 1686, defined as the product of and the square of (mv²), which served as a precursor to and highlighted the conservation of a associated with motion, contrasting with earlier momentum-based views like Descartes' quantité de mouvement. Leibniz's work laid groundwork for distinguishing dynamic forces from positional ones, though the full separation into kinetic and potential forms emerged later. In the , the concept advanced through applications in , where introduced the function in the 1780s within his Mécanique Céleste, enabling the calculation of gravitational forces as the gradient of a rather than direct integration of inverse-square laws. This approach simplified analyses of planetary perturbations and stability. Complementing Laplace, developed a general of potential in 1839–1840, formalizing the for attractive and repulsive forces, including , and introducing key mathematical tools like for geomagnetic and gravitational computations. These developments marked a shift toward using potentials to model conservative fields efficiently. The saw the formalization of as a distinct term and its integration into broader conservation principles. Scottish engineer William John Macquorn Rankine coined the phrase "potential energy" in 1853, explicitly denoting the stored due to position or configuration in a force field, such as gravitational or elastic systems, to distinguish it from "actual energy" (kinetic). advanced this in his 1847 memoir "On the Conservation of Force," demonstrating that potential energy, alongside kinetic, remains invariant across mechanical, electrical, and chemical transformations, unifying diverse phenomena under a single . William Thomson (later ) further popularized these ideas in the , collaborating on thermodynamic applications and emphasizing potential's role in dissipation and equilibrium. This evolution reflected a pivotal post-1800 from momentum-centric to , influenced by experiments like those of James Joule on heat-work equivalence, enabling solutions to complex problems such as planetary motion through potential-based variational methods rather than exhaustive force integrations. The concept's adoption revolutionized physics by providing a for non-contact interactions, facilitating advancements in and .

Theoretical Framework

Conservative Forces

A conservative force is a force for which the work done by the force on an object as it moves between two points depends only on the initial and final positions of the object, and not on the specific path taken between those points. This path independence implies that the work done around any closed path is zero. In mathematical terms, a vector force field \mathbf{F} is conservative if its curl vanishes everywhere in a simply connected domain, that is, \nabla \times \mathbf{F} = \mathbf{0}. This condition ensures that the line integral \int_C \mathbf{F} \cdot d\mathbf{r} along any path C connecting two points is independent of the path chosen and equals the difference in a scalar function evaluated at the endpoints. Consequently, such a force can be expressed as the negative gradient of a scalar potential function U, so \mathbf{F} = -\nabla U. Examples of conservative forces include the gravitational force acting on masses, the electrostatic force between charged particles, and the restoring force exerted by an ideal spring. In contrast, non-conservative forces, such as , dissipate in a manner that depends on the path traversed, leading to work that varies with the details of the motion. The defining properties of conservative forces form the foundation for potential energy, as they permit the storage of in a reversible form without net loss during motion. Specifically, the work done by a conservative force equals the negative change in the associated potential energy, \Delta U = -\int \mathbf{F} \cdot d\mathbf{r}, which conserves the total in isolated systems governed solely by such forces. This reversibility distinguishes conservative forces as prerequisites for defining potential energy functions in .

Relation to Work

The relationship between potential energy and work arises in the context of conservative forces, which are defined such that the work they perform depends only on the initial and final positions of an object, independent of the path taken. For such forces, the change in potential energy \Delta U of a system is equal to the negative of the work W done by the conservative force: \Delta U = -W. This relation implies that when a conservative force does positive work on an object, the system's potential energy decreases, converting it into other forms such as . In conservative systems, this connection adapts the work-energy theorem, which generally states that the net work done on an object equals its change in . For forces that are exclusively conservative, the work done by these forces W_\text{conservative} equals the negative change in potential energy: W_\text{conservative} = -\Delta U. Consequently, the total mechanical energy, defined as the sum of K and potential energy U, remains constant: \Delta K + \Delta U = 0, provided no non-conservative forces (such as ) perform work. This holds because the work by conservative forces merely redistributes energy between kinetic and potential forms without dissipation. The vector formulation of this relation expresses the work done by a conservative \mathbf{F} along a path as the W = \int \mathbf{F} \cdot d\mathbf{r}. Since \mathbf{F} = -\nabla U for a conservative , where \nabla U is the of the potential energy function, the becomes W = -\int \nabla U \cdot d\mathbf{r} = U_i - U_f = -\Delta U, with U_i and U_f denoting initial and final potential energies. This equality confirms the path independence of the work, as the of a depends solely on the endpoints. These relations enable the prediction of an object's motion in conservative fields without computing the full or integrating forces at every point, relying instead on differences between positions. For instance, in systems like gravitational or electrostatic fields, the path-independent nature allows straightforward analysis of transfers, facilitating solutions to complex dynamics through conservation principles alone.

Computing Potential Energy

The potential energy U(\mathbf{r}) for a conservative \mathbf{F}(\mathbf{r}) is defined such that the work done by the force along any path from a reference point \mathbf{r}_0 to \mathbf{r} equals the negative change in potential energy: W = U(\mathbf{r}_0) - U(\mathbf{r}). This leads to the general computational method via the path-independent U(\mathbf{r}) = -\int_{\mathbf{r}_0}^{\mathbf{r}} \mathbf{F} \cdot d\mathbf{l}, where the integral can be evaluated analytically if \mathbf{F} is known explicitly, and U(\mathbf{r}_0) is set to an arbitrary constant, often zero. The choice of reference point \mathbf{r}_0 determines this constant and is arbitrary, as detailed in the section on Choice of Reference Level. For systems involving multiple independent conservative forces \mathbf{F}_i, the superposition principle applies because the total force is the vector sum \mathbf{F} = \sum_i \mathbf{F}_i, and the total work is the sum of individual works. Consequently, the total potential energy is the sum of the individual potentials: U(\mathbf{r}) = \sum_i U_i(\mathbf{r}). To verify a computed potential or derive the force from a given U, the relationship \mathbf{F} = -\nabla U is used, where the components are the negative partial derivatives: F_x = -\partial U / \partial x, F_y = -\partial U / \partial y, and F_z = -\partial U / \partial z. This assumes the force field is known from specific laws, as covered in later sections on gravitational and potentials. In cases of complex force fields where analytical integration is impractical, numerical methods are employed to approximate the line integral or solve the gradient equation \nabla U = -\mathbf{F}, such as through discretization techniques like finite differences on a grid.

Reference and Measurement

Choice of Reference Level

The potential energy U of a system is inherently defined relative to an arbitrary zero point, meaning its absolute value lacks unique physical meaning; instead, only the change in potential energy \Delta U between states is observable and corresponds to the negative of the work done by conservative forces along a path. This arbitrariness arises because the definition of potential energy involves an indefinite integral of the force, which introduces an additive constant that can be chosen freely without affecting the laws of physics, such as energy conservation. Consequently, calculations focus on differences rather than absolute values to determine energy transfers or mechanical equilibria. Common choices for the reference level are selected for mathematical convenience or practical . For inverse-distance potentials like those in gravitational or electrostatic interactions, the zero of potential energy is often set at infinite separation, where interactions vanish, simplifying the analysis of bound orbits or processes. In contrast, for near-surface gravitational scenarios, such as objects on , the reference is typically the ground level or the minimum height in the system, allowing straightforward computation of changes during motion without needing to account for planetary scales. These conventions ensure that the chosen zero aligns with the problem's context while preserving the invariance of \Delta U. Certain potentials, such as the Newtonian gravitational one, are unbounded below, approaching negative as the separation between bodies approaches , which might suggest ill-defined energies for close approaches. However, since physical predictions rely solely on finite differences \Delta U, this unboundedness poses no issue; the energy conservation principle remains robust as long as the reference is consistently applied, even for highly negative absolute values in bound systems. In multi-body systems, where potential energy is the sum of pairwise contributions, maintaining a uniform reference level across all components is essential to avoid inconsistencies in the total . This often involves adopting a global zero, such as the configuration where all inter-body distances tend to , ensuring that the collective \Delta U accurately reflects interactions without artificial offsets between subsystems.

Units and Dimensions

The unit of potential energy is the joule (J), defined as 1 J = 1 kg·m²/s², which is the same unit used for work and . This equivalence arises because potential energy represents stored work that can be converted into , ensuring dimensional consistency in energy transformations. The dimensional formula for potential energy is [M L^2 T^{-2}], matching that of all forms of and underscoring the principle of where total remains invariant. In various physical systems, potential energy scales differently with mass and distance parameters. For , it scales linearly with the object's mass and the vertical distance from a reference level, reflecting the direct proportionality to positional changes in a uniform . For potential in deformable systems like springs, it scales quadratically with from , capturing the nonlinear storage in strained configurations. At atomic and subatomic scales, where joules yield impractically small values, potential energies are commonly expressed in electronvolts (), with 1 eV = 1.602 × 10^{-19} J, facilitating analysis in quantum and contexts. While absolute potential energy values depend on the selected reference level, measurable differences between states are independent of this choice and quantified in these units.

Gravitational Potential Energy

Near-Earth Approximation

In the near-Earth approximation, gravitational potential energy is modeled as a linear function of height, suitable for objects close to Earth's surface where the gravitational field can be treated as uniform. This simplification arises from the constant gravitational force F = mg, where m is the mass of the object and g is the acceleration due to gravity, approximately $9.8 \, \mathrm{m/s^2}. The potential energy U relative to a reference height is then given by U = m g h, where h is the height above the reference level. This formula represents the work done to lift the object against gravity in a constant field, and only differences in U (i.e., \Delta U = m g \Delta h) have physical significance for energy conservation in motion. The key assumptions underlying this approximation include a gravitational field, which holds when the h is much smaller than Earth's radius (about 6371 km), ensuring negligible variation in g over the distance. It derives directly from integrating the constant F = mg over height, yielding a linear potential without accounting for the inverse-square nature of at larger scales. This model is valid for everyday terrestrial applications, such as calculating changes in vertical motion near the surface. Representative examples illustrate its utility. For a book of mass 1 kg placed on a shelf 2 m above the floor, the potential energy stored is U = (1 \, \mathrm{kg})(9.8 \, \mathrm{m/s^2})(2 \, \mathrm{m}) = 19.6 \, \mathrm{J}, which converts to kinetic energy if the book falls. In larger-scale applications, hydroelectric dams harness this energy; water held at a height behind a dam, such as 50 m, releases gravitational potential energy as it flows downward, driving turbines to generate electricity—for instance, for a reservoir mass of $10^{12} \, \mathrm{kg}, the available energy difference is on the order of $5 \times 10^{14} \, \mathrm{J}. These cases highlight how \Delta U drives mechanical work in systems like pendulums or falling objects. This approximation breaks down for large heights, such as in orbital mechanics or high-altitude flights, where the varying gravitational field requires the more general two-body formula for accuracy. It remains indispensable, however, for engineering and introductory analyses on Earth-bound scales.

General Two-Body Formula

The gravitational potential energy U for two point masses M and m separated by a distance r is given by the formula U = -\frac{G M m}{r}, where G is the gravitational constant. This expression arises from Newton's law of universal gravitation and represents the work required to separate the masses from their current positions to infinite separation. The negative sign in the indicates that is attractive, as the potential energy decreases (becomes more negative) when the masses approach each other. By , U = 0 when r \to \infty, reflecting that no work is needed to assemble the system from infinitely separated masses. The $1/r dependence highlights how the potential weakens inversely with separation, distinguishing it from linear approximations valid only over small distances. For extended bodies that are spherically symmetric, such as uniform spheres, the gravitational potential energy can be approximated by treating each body as a point mass located at its center of mass, provided the separation r is much larger than the bodies' radii. This equivalence follows from Newton's shell theorem, which shows that the gravitational field outside a spherical mass distribution matches that of a point mass at the center. In a multi-body system consisting of N point masses, the total gravitational potential energy is the sum of the pairwise interactions: U = -\sum_{i < j} \frac{G m_i m_j}{r_{ij}}, where r_{ij} is the distance between masses i and j. This pairwise summation assumes no higher-order interactions beyond two-body terms, making it exact for Newtonian gravity in the absence of relativistic effects. The derivation of the two-body formula via integration over mass distributions is detailed in the section on Derivation for Gravitational Potential.

Derivation for Gravitational Potential

The gravitational potential energy for two point masses arises from , which states that the force \mathbf{F} between masses M and m separated by a r is given by \mathbf{F} = -\frac{[G](/page/Gravitational_constant) M m}{r^2} \hat{r}, where G is the and \hat{r} is the unit vector pointing from M to m. To derive the potential energy U(r), recall that for a conservative , the change in potential energy is the negative of the work done by the , \Delta U = -W, with the work computed as the \int \mathbf{F} \cdot d\mathbf{l}. Choosing the reference point where U(\infty) = 0, the potential energy at separation r is \begin{aligned} U(r) &= -\int_{\infty}^{r} \mathbf{F} \cdot d\mathbf{l} \\ &= -\int_{\infty}^{r} \left( -\frac{G M m}{s^2} \right) ds \\ &= G M m \int_{\infty}^{r} \frac{1}{s^2} ds \\ &= G M m \left[ -\frac{1}{s} \right]_{\infty}^{r} \\ &= -\frac{G M m}{r}, \end{aligned} where the integral is taken along a radial path, with s as the dummy variable for separation distance. This derivation relies on the path independence of the line integral, which holds because the gravitational force is conservative; its curl vanishes (\nabla \times \mathbf{F} = 0) for the inverse-square law, ensuring the work depends only on initial and final positions. For the two-body problem, this expression U(r) = -\frac{G M m}{r} fully captures the gravitational potential energy, though generalizations exist for systems with variable mass distributions or non-point sources by integrating over volume elements.

Sign Convention and Applications

In gravitational potential energy, the sign convention defines the potential U as negative for bound systems, where the reference point of zero potential is conventionally set at infinity. This choice arises from the integration of the gravitational force, resulting in U = -\frac{GMm}{r}, which ensures that work done by gravity is positive as objects move closer together. For such systems, the escape energy required to reach infinity with zero kinetic energy equals the negative of the potential energy, -U, representing the minimum energy input needed to unbind the system. The negative sign implies that gravitationally bound states, such as orbiting stars, possess negative total , preventing without external input. In these configurations, the total energy E = K + U is negative, where kinetic energy K is positive but insufficient to overcome the deeper negative . Parabolic orbits, marking the boundary between bound and unbound trajectories, have zero total energy, allowing objects to coast to with vanishing speed. This convention highlights the of bound systems like planetary orbits, where perturbations must supply energy exceeding -E to achieve . In , the negative potential underpins the , which conserves energy to relate an object's speed at any point to its orbital parameters, enabling predictions of satellite trajectories without recomputing forces at each step. For black holes, the extremely negative potential near the event horizon traps and , with Newtonian approximations yielding potentials orders of magnitude deeper than 's, illustrating binding energies that approach infinite depth. Tidal energy harnesses gradients in this potential between , , and Sun, converting differential gravitational pulls into kinetic flows that drive ocean currents, powering renewable systems with global potential exceeding 3 terawatts. Modern applications leverage this convention for . In GPS systems, satellite clocks run faster by about 45 microseconds daily due to weaker negative in compared to Earth's surface, necessitating relativistic corrections to maintain positional accuracy within meters. launches account for in staging designs; for instance, the theoretical minimum for reaching is roughly 32 megajoules per kilogram to overcome the planet's binding potential, while full to interplanetary demands an additional approximately 30 megajoules per kilogram, guiding fuel efficiencies in missions like those to Mars.

Elastic Potential Energy

Linear Spring Model

The linear spring model provides a foundational description of elastic potential energy in systems where materials deform proportionally under applied forces, such as in ideal springs. This model assumes the material behaves linearly, with the restoring directly opposing the deformation and proportional to its magnitude, applicable primarily to small extensions or compressions that do not exceed the limit of the . Central to this model is , which states that the force F exerted by the spring is F = -kx, where x is the from the position and k is the spring constant, a measure of the spring's with units of newtons per meter (N/m). The negative sign indicates the restorative nature of the force, directed toward the . This linear relationship holds under the of small deformations, where the spring's response remains proportional and reversible without or plastic yielding. The elastic potential energy U stored in the deformed spring is given by the quadratic formula U = \frac{1}{2} k x^2, which quantifies the available to do work upon release. This expression derives from the work-energy principle, where U equals the work required to deform the from , computed as the of the conservative : U(x) = -\int_0^x F \, dx = \int_0^x k s \, ds = \frac{1}{2} k x^2. The assumes a quasi-static process with negligible during deformation, deferring detailed treatment of dynamic effects to broader analyses of oscillatory systems. In practical applications, the linear spring model underpins the analysis of - systems undergoing (SHM), where a m attached to the oscillates with the total conserved as E = \frac{1}{2} k A^2, with A as the ; at maximum , all is potential (U = E), converting fully to kinetic at equilibrium. Similarly, for small angular s (\theta \ll 1 ), the simple approximates SHM akin to a , with an effective restoring leading to an equivalent potential energy form that mirrors the quadratic dependence on . These examples illustrate the model's utility in predicting periodic behaviors in mechanical oscillators, from components to geophysical phenomena.

General Elastic Deformations

In , the elastic potential energy stored in a deformable solid due to general deformations is given by the integral of the over the volume of the material: U = \frac{1}{2} \int_V \sigma_{ij} \epsilon_{ij} \, dV where \sigma_{ij} is the tensor, \epsilon_{ij} is the infinitesimal tensor, and the is performed over the volume V of the body. This expression assumes small deformations where the material response is linear, meaning is proportional to via the . For isotropic materials, the relationship between and is characterized by two independent constants: E, which quantifies resistance to uniaxial or , and \nu, which describes the ratio of transverse to axial under uniaxial loading. In bulk materials, the effective spring constant k for a , such as a of L and cross-sectional area A, relates to as k = E A / L, illustrating how material stiffness scales with geometry. enters the full constitutive relations, affecting the , for example, u = \frac{1}{2E} [\sigma_{xx}^2 + \sigma_{yy}^2 + \sigma_{zz}^2 - 2\nu (\sigma_{xx}\sigma_{yy} + \sigma_{yy}\sigma_{zz} + \sigma_{zz}\sigma_{xx})] + \frac{1}{2G} (\sigma_{xy}^2 + \sigma_{yz}^2 + \sigma_{zx}^2), where G = E / [2(1 + \nu)] is the . This framework applies to various deformation modes in . In beam bending, the strain energy arises primarily from normal stresses varying linearly across the cross-section, leading to U = \int_0^L \frac{M(x)^2}{2 E I} \, dx, where M(x) is the and I is moment of area; this captures the energy stored during deflection under transverse loads. For torsion in cylindrical shafts, the energy is due to stresses, expressed as U = \int_0^L \frac{T(x)^2}{2 G J} \, dx, with T(x) the and J the polar moment of inertia, relevant for twisted structural components. At the atomic scale, vibrations in crystals embody potential energy through interatomic forces, quantized as phonons—collective modes that contribute to and properties without net macroscopic deformation. For larger strains beyond the linear regime, such as in elastomers, nonlinear extensions replace the quadratic strain energy density with hyperelastic models that depend on deformation invariants. , for instance, is often modeled using the neo-Hookean form W = \frac{\mu}{2} (I_1 - 3), where W is the density, \mu is a parameter, and I_1 is the first invariant of the right Cauchy-Green deformation tensor; this accounts for entropic restoring forces in polymer networks under finite s up to several hundred percent. More general models, like the Ogden form, extend this by incorporating multiple terms to fit experimental stress- data across , , and . The linear serves as a special case for these descriptions in one-dimensional, small-deformation limits.

Electrostatic Potential Energy

Between Point Charges

The electrostatic potential energy between two point charges quantifies the work required to assemble the charges from separation to a finite , under the of their mutual interaction. This energy is a fundamental concept in , applicable to discrete charge configurations where continuous distributions are not considered. For two static, point-like charges q_1 and q_2 in , separated by distance r, the potential energy U is expressed as U = \frac{1}{4\pi\epsilon_0} \frac{q_1 q_2}{r}, where \epsilon_0 is the permittivity of free space. This formula assumes the charges are idealized points with no spatial extent and remain at rest, neglecting relativistic or quantum effects. The potential is defined relative to zero at infinite separation, ensuring U \to 0 as r \to \infty. The sign of U reflects the nature of the interaction: positive for like charges (repulsive , requiring external work to bring them closer) and negative for opposite charges (attractive , with the system doing work as charges approach). This highlights the of opposite-charge configurations at finite distances. The of this expression from the integral of is outlined in the dedicated section on electrostatic potential energy origins. A representative example occurs in ion pairs, such as the ^+ and ^- in ionic crystals, where the attractive potential energy (negative U) contributes dominantly to the lattice binding, on the order of several electronvolts per pair despite lattice summation effects. In the context of capacitor charging, the total stored arises from pairwise interactions between accumulated charges on the plates, approximated as point charges for small separations, yielding stored energies proportional to q^2 / (2C) for capacitance C. This energy is fundamentally stored in the pervading the space between the charges. For assemblies exceeding two charges, the total electrostatic potential energy is the scalar over all unique pairwise contributions, as explored in the subsequent section on systems of multiple charges.

Systems of Multiple Charges

In systems consisting of multiple point charges, the total electrostatic potential energy is given by the over all unique pairs of charges to account for their interactions without double-counting: U = \frac{1}{2} \sum_{i \neq j} \frac{1}{4\pi \epsilon_0} \frac{q_i q_j}{r_{ij}}, where q_i and q_j are the charges, r_{ij} is the between them, and the factor of $1/2 ensures each pair is counted once. This expression extends the pairwise potential energy for two charges by aggregating contributions from all pairs in the assembly. A notable aspect is the self-energy term for an individual point charge, which arises from the of the charge with its own and formally diverges to infinity in the classical point-charge model, requiring regularization techniques such as finite or quantum mechanical treatments for physical applications. In practice, this infinite is often excluded or renormalized when computing the total of finite systems, focusing instead on energies. Applications of this formulation include calculating the electrostatic contributions to the energies in molecular ions, such as H₂⁺, where the potential energy from nuclear and electronic charge distributions stabilizes the ion. For continuous charge distributions, the total electrostatic potential energy can be expressed as an integral over the throughout : U = \frac{\epsilon_0}{2} \int E^2 \, dV, where the integration covers all , and E is the magnitude of the produced by the distribution. This field-based form is equivalent to the charge-potential U = \frac{1}{2} \int \rho \phi \, dV, with \rho as and \phi as , and proves useful for complex geometries. In dielectric media, the presence of polarizable materials modifies the energy expression, replacing the vacuum permittivity \epsilon_0 with the material's permittivity \epsilon = \kappa \epsilon_0, where \kappa is the dielectric constant, yielding an energy density of \frac{1}{2} \mathbf{E} \cdot \mathbf{D} = \frac{\epsilon}{2} E^2 for linear isotropic , with \mathbf{D} as the . This adjustment accounts for the reduced field strength and energy storage in materials like capacitors filled with insulators, enhancing compared to .

Derivation from Coulomb's Law

The electrostatic force between two stationary point charges q_1 and q_2 separated by a distance r is given by Coulomb's law: \vec{F} = \frac{1}{4\pi\epsilon_0} \frac{q_1 q_2}{r^2} \hat{r}, where \epsilon_0 is the vacuum permittivity, \hat{r} is the unit vector pointing from q_1 to q_2, and the force is repulsive if q_1 and q_2 have the same sign. This force arises from experimental observations and is fundamental to electrostatic interactions. The force is conservative, meaning the work done by it to move a charge between two points depends only on the endpoints and not on the path taken, due to its inverse-square dependence and central nature (with zero curl in ). This path independence allows the definition of a energy function U(r), analogous to the gravitational potential energy between masses, where the gravitational force also follows an but is always attractive. For static charges, the potential energy is associated solely with their positions, without time-varying fields. To derive the electrostatic potential energy, consider fixing q_1 at the and bringing q_2 from (where U = 0) to distance r. The change in potential energy is the negative of the work done by the electrostatic : U(r) = -\int_{\infty}^{r} \vec{F} \cdot d\vec{l}. Along the radial path, d\vec{l} = dr' \hat{r} (with dr' negative inward, but handled by limits), and \vec{F} \cdot d\vec{l} = \frac{1}{4\pi\epsilon_0} \frac{q_1 q_2}{r'^2} dr', yielding U(r) = -\int_{\infty}^{r} \frac{1}{4\pi\epsilon_0} \frac{q_1 q_2}{r'^2} dr' = \frac{1}{4\pi\epsilon_0} \frac{q_1 q_2}{r}. This expression is positive for like charges (requiring work against repulsion) and negative for opposite charges (releasing energy). For static configurations of point charges, this pairwise form generalizes directly, though multi-charge systems require summing over interactions while accounting for assembly order to avoid double-counting.

Magnetic Potential Energy

Dipoles in Magnetic Fields

The magnetic potential energy of a in a arises from the interaction between the dipole's and the external field, analogous to gravitational or electrostatic potentials but specific to magnetic configurations. For a magnetic with moment \vec{\mu}, the potential energy U in a uniform \vec{B} is given by U = -\vec{\mu} \cdot \vec{B} = -\mu B \cos\theta, where \theta is the angle between \vec{\mu} and \vec{B}. This expression shows that the energy is minimized when the dipole is aligned parallel to the field (\theta = 0^\circ), reaching a maximum when anti-aligned (\theta = 180^\circ). The \vec{\tau} on the , which tends to align it with , can be derived from the negative of the potential energy: \vec{\tau} = -\nabla U. In a uniform field, this simplifies to \vec{\tau} = \vec{\mu} \times \vec{B}, with magnitude \tau = \mu B \sin\theta. This relation highlights the rotational force acting on the until is achieved at minimum potential energy. A classic example is the needle, a bar with \vec{\mu} that aligns with \vec{B} to minimize its potential energy, pointing north. In paramagnetic materials, atomic or molecular dipoles experience a similar tendency in an applied , though agitation limits full , leading to a net proportional to the field strength. Another application occurs in (NMR), where the potential energy difference between aligned and anti-aligned nuclear spins in a strong \Delta U = 2\mu B determines the energy splitting and frequency for spectroscopic analysis. This formulation assumes quasi-static , where the field varies slowly compared to the system's response time, and neglects effects, valid when the size is much smaller than the wavelength of any associated .

Current Loops and Conductors

In -carrying conductors, such as loops or solenoids, magnetic potential energy arises from the interaction between the and the it generates, primarily through effects. For a single circuit, the magnetic potential energy stored is given by U = \frac{1}{2} L I^2, where L is the self- of the circuit and I is the steady flowing through it. This energy represents the work done to establish the against the opposing self-induced , and it is conserved in the absence of dissipation. Self-inductance L quantifies how effectively the circuit's own links with itself, as in a where L depends on the number of turns, , and surrounding medium. For example, in practical inductors used in electronic circuits, this stored enables functions like filtering signals or storing energy in switched-mode power supplies, with typical values of L ranging from microhenries to henries. The energy formula holds under the quasi-static approximation, where currents vary slowly enough that radiative losses and retardation effects are negligible, allowing the to be treated as effectively conservative. When two circuits interact, mutual inductance introduces an additional term to the total magnetic . For two circuits carrying currents I_1 and I_2, the interaction energy is U = M I_1 I_2, where M is the mutual measuring the through one circuit due to the current in the other. This term can be positive or negative depending on the relative of the currents, reflecting whether the fields aid or oppose each other. In transformers, mutual inductance M couples primary and secondary windings to transfer efficiently between , with the total stored energy including both self and mutual contributions: U = \frac{1}{2} L_1 I_1^2 + \frac{1}{2} L_2 I_2^2 + M I_1 I_2. The inductance-based energy expressions are equivalent to the energy stored in the itself. The total is U = \frac{\mu_0}{2} \int B^2 \, dV, integrated over all space, where B is the and \mu_0 is the permeability of free space. For a or , this integral yields the same \frac{1}{2} L I^2 when the field is confined, confirming that the energy resides in the field rather than the conductor. This equivalence underscores the macroscopic electromagnetic nature of the potential energy in such systems, distinct from microscopic orientations.

Chemical Potential Energy

Molecular Bond Energies

Molecular bond energies represent the potential energy stored in the chemical bonds of molecules, arising primarily from electrostatic interactions between nuclei and electrons, augmented by quantum mechanical effects that stabilize shared electron pairs in covalent bonds and in ionic bonds. The (BDE) quantifies this, defined as the energy required to break a specific into fragments under conditions, reflecting the depth of the potential energy well associated with that bond. For instance, the H–H in the dihydrogen molecule has a BDE of approximately 436 kJ/mol, corresponding to the energy minimum on its (PES) at the of about 0.74 . This PES describes the total molecular energy as a of nuclear coordinates, with the occurring at a local minimum where the forces on the nuclei vanish. The quantum mechanical foundation of these bond energies relies on the Born-Oppenheimer approximation, which separates electronic and nuclear motions due to the much larger mass of nuclei, allowing the electronic energy to be computed as a function of fixed nuclear positions to generate the PES. Within this framework, methods like Hartree-Fock theory solve the electronic by assuming a single wavefunction, providing approximate bond energies, while (DFT) improves accuracy by incorporating electron correlation effects through exchange-correlation functionals, often yielding bond energies within chemical accuracy (about 4 kJ/mol) for many systems. These computational approaches reveal how quantum —arising from the indistinguishability of electrons—lowers the energy below classical electrostatic predictions, enabling stable covalent bonding. In practical applications, molecular bond energies serve as the basis for chemical potential energy storage in fuels like hydrocarbons, where high-energy C–H and C–C bonds store derived from their formation, and in batteries, such as lithium-ion systems, where intercalation involves bond rearrangements that store and release upon charge-discharge cycles. The release of this stored occurs when bonds are broken or reformed during chemical processes, converting potential energy into other forms.

Role in Chemical Reactions

In chemical reactions, the potential energy landscape is described by the (PES), a multidimensional map of potential energy as a function of atomic positions for the reacting . Along the —the minimum-energy pathway connecting reactants to products—the PES features a at the , where the potential energy reaches a local maximum. The activation energy E_a corresponds to the difference in potential energy \Delta U between the reactants and this transition state, determining the energy barrier that must be overcome for the reaction to proceed. Reactions are characterized as endothermic or exothermic based on the sign of the potential change \Delta U: positive for endothermic processes where products have higher potential than reactants, and negative for exothermic ones where is released. While \Delta U provides insight into the energetic driving force, the overall feasibility and direction of a under constant and pressure are governed by the change, given by \Delta G = \Delta U - T \Delta S + P \Delta V, where T is , \Delta S is the change, and \Delta V is the volume change; a negative \Delta G indicates a spontaneous . Representative examples illustrate these concepts: combustion of hydrocarbons, such as with oxygen, is a highly that releases substantial potential energy stored in molecular bonds, converting it to and with \Delta U on the order of hundreds of kJ/mol. In , energy drives an in , storing potential energy by forming glucose from and , effectively increasing \Delta U. Catalysts, such as enzymes in biological systems, accelerate by stabilizing the and lowering the barrier through an alternative pathway on the PES, without altering the overall \Delta U. These potential energy changes, including differences arising from molecular bond energies, are experimentally measured using , which detects at constant (bomb calorimetry) or volume to quantify enthalpies and infer bond dissociation energies with precisions often better than 1 kJ/mol.

Nuclear Potential Energy

Strong Nuclear Force Potential

The , responsible for binding protons and neutrons within atomic nuclei, is modeled at low energies by short-range potentials that capture its attractive nature over distances of approximately 1 femtometer (fm). This force dominates over electromagnetic repulsion at nuclear scales, enabling stable nuclear structures. The foundational model is the Yukawa potential, proposed by Hideki Yukawa in 1935, which describes the interaction between nucleons as an exchange of massive particles, later identified as pions. The potential takes the form U(r) \approx -\frac{g^2}{4\pi} \frac{e^{-m r}}{r} + \text{repulsive core at small } r, where g is the coupling constant, m is the pion mass (approximately 140 MeV/c²), and r is the separation distance; the exponential decay ensures a finite range unlike the infinite-range Coulomb potential. At distances around 1 fm, the potential is strongly attractive, while a hard repulsive core emerges for r \lesssim 0.5 fm due to overlapping quark wavefunctions and Pauli exclusion effects, preventing nucleons from collapsing. Key properties include mediation by virtual pion exchange, which imparts a spin- and isospin-dependent character to the force. The strong force exhibits charge independence, acting equally between proton-proton, neutron-neutron, and proton-neutron pairs, as the underlying quark-level interactions are flavor-blind. For mean-field approximations in nuclear structure calculations, phenomenological models like the Woods-Saxon potential are employed, given by U(r) = -U_0 \left[1 + \exp\left(\frac{r - R}{a}\right)\right]^{-1}, where U_0 is the depth (around MeV), R approximates the , and a sets the surface diffuseness (about 0.5-0.7 fm); this form effectively averages interactions across the . The successfully explains the of the deuteron, the simplest (proton-neutron pair), with a of 2.224 MeV arising from the delicate balance of attraction and repulsion. Overall, the potential vanishes beyond 2-3 fm, confining its effects to dimensions and dropping to negligible values outside. This short underpins stability while allowing for the observed in multi-nucleon systems.

Nuclear Binding Energies

Nuclear binding energy is the minimum energy required to disassemble an atomic nucleus into its constituent protons and neutrons, representing the stability imparted by the strong nuclear force. This energy originates from the attractive potential between nucleons, which overcomes the repulsive Coulomb force between protons, resulting in a net negative potential energy for the bound system. The binding energy quantifies the depth of this potential well; for instance, in light nuclei like helium-4, the total binding energy is approximately 28.3 MeV, corresponding to about 7.07 MeV per nucleon. The binding energy is calculated using the mass defect principle, derived from Einstein's mass-energy equivalence E = mc^2. The mass defect \Delta m is the difference between the mass of the isolated nucleons and the mass of the nucleus: \Delta m = Z m_p + N m_n - M(A, Z) where Z is the atomic number, N = A - Z is the neutron number, A is the mass number, m_p and m_n are the masses of the proton and neutron, and M(A, Z) is the nuclear mass. The total binding energy B is then B = \Delta m c^2. This approach reveals that bound nuclei have less mass than their separated parts, with the "missing" mass converted to binding energy. For deuterium (^2_1H), the binding energy is 2.224 MeV, illustrating the strong force's role in overcoming proton-proton repulsion. To model binding energies across nuclei, the approximates B(A, Z) by balancing nuclear and electromagnetic contributions: B(A, Z) = a_v A - a_s A^{2/3} - a_c \frac{Z(Z-1)}{A^{1/3}} - a_a \frac{(A - 2Z)^2}{A} \pm \delta Here, a_v \approx 15.5 MeV accounts for the volume saturation of the strong force potential; a_s \approx 16.8 MeV for surface effects reducing attraction at the boundary; a_c \approx 0.72 MeV for repulsion; a_a \approx 23.3 MeV for due to neutron-proton imbalance; and \delta is a pairing term favoring even-even nuclei. This formula, rooted in the liquid drop model, predicts binding energies with good accuracy for medium to heavy nuclei. The per B/A peaks at around 8.8 MeV for and , indicating maximum stability near A \approx 56. For lighter nuclei (A < 56), B/A increases with A due to growing strong force contributions, enabling energy release in fusion; for heavier nuclei, it decreases owing to dominant Coulomb repulsion, favoring fission. This curve explains stellar nucleosynthesis and nuclear power: fusion of hydrogen to helium releases ~7 MeV per nucleon, while uranium-235 fission yields ~200 MeV per event.

References

  1. [1]
    Potential Energy - HyperPhysics
    Potential energy is energy from position or configuration, like gravitational, electric, magnetic, or elastic energy. It's the work needed to move an object ...Missing: authoritative | Show results with:authoritative
  2. [2]
    Work, Kinetic Energy, and Potential Energy - Mechanics Map
    Potential energy can come in many forms, but the two we will discuss here are gravitational potential energy, and elastic potential energy. As the names imply, ...
  3. [3]
    None
    ### Summary of Potential Energy from Lecture Note
  4. [4]
    8.1 Potential Energy of a System – University Physics Volume 1
    This property allows us to define a different kind of energy for the system than its kinetic energy, which is called potential energy.<|control11|><|separator|>
  5. [5]
    Potential Energy - The Physics Classroom
    Potential energy is the stored energy of position, relative to a zero position. It includes gravitational and elastic forms.Missing: authoritative sources
  6. [6]
    Potential Energy – Physics 131: What Is Physics? - Open Books
    Experiments show that these three factors all contribute equally: mass, height, and gravitational strength are all of equal importance to the gravitational ...
  7. [7]
    13 Work and Potential Energy (A) - Feynman Lectures
    The simplest example of the conservation of energy is a vertically falling object, one that moves only in a vertical direction.
  8. [8]
    The Master Equation for Energy Problems - Physics
    Potential energy is energy associated with position. Usually we view potential energy as energy that is stored temporarily. One type of potential energy is ...
  9. [9]
    Thermodynamics: Kinetic and Potential Energy
    Potential energy is energy an object has because of its position relative to some other object. When you stand at ...
  10. [10]
    Conservation of energy - Physics
    Oct 13, 1999 · Conservation of mechanical energy means the total mechanical energy (potential + kinetic) remains constant with conservative forces, like ...
  11. [11]
    8. THE CONSERVATION OF ENERGY - Home Page of Frank LH Wolfs
    Conservation of energy tells us that the total energy of the system is conserved, and in this case, the sum of kinetic and potential energy must be constant.
  12. [12]
    The concept of energy and its early historical development
    The concept of energy and its early historical development. Published ... Lindsay, R.B. The concept of energy and its early historical development.
  13. [13]
    Pierre-Simon Laplace (1749 - 1827) - Biography - MacTutor
    Pierre-Simon Laplace proved the stability of the solar system. In analysis Laplace introduced the potential function and Laplace coefficients. He also put ...
  14. [14]
    Carl Friedrich Gauss (1777 - 1855) - Biography - MacTutor
    At the age of seven, Carl Friedrich Gauss started elementary school, and his potential was noticed almost immediately.Missing: date | Show results with:date
  15. [15]
    What is potential energy? - IOPscience
    Feb 25, 2003 · W J M Rankine coined the term 'potential energy' 150 years ago. This paper examines why he introduced it, the evolution of its ...
  16. [16]
    Hermann von Helmholtz - Stanford Encyclopedia of Philosophy
    Feb 18, 2008 · On the 23rd of July in 1847, Helmholtz gave an address, “The Conservation of Force,” at the Physical Society. “Force” [Kraft], as Helmholtz uses ...Biographical note and... · Theory of Perception · Conservation laws... · Bibliography
  17. [17]
    7.4 Conservative Forces and Potential Energy - UCF Pressbooks
    A conservative force is one for which work done by or against it depends only on the starting and ending points of a motion and not on the path taken.
  18. [18]
    8.2 Conservative and Non-Conservative Forces - UCF Pressbooks
    A conservative force is one for which the work done is independent of path. Equivalently, a force is conservative if the work done over any closed path is zero.
  19. [19]
    [PDF] U ○ Section 4.3: Force as the gradient of potential energy ...
    Theorem. For a conservative force F,. ∇ × F = 0 . Proof: The curl of a gradient is always 0... Since F can be written as a gradient,. i.e., F = − ∇U ,. ∇ × F = ...
  20. [20]
    [PDF] Conservative fields and potential functions. (Sect. 16.3)
    F · dr = f (r1) − f (r0). F · dr is path independent. The line integral of conservative fields. F · dr = f (r1) − f (r0).
  21. [21]
    [PDF] PHYS 419: Classical Mechanics Lecture Notes
    Theorem 2a: For any conservative force field F(r) there exists a scalar function U(r) such that F(r) = −VU(r). The function U(r) is called the potential energy.<|control11|><|separator|>
  22. [22]
    [PDF] Conservative and Non-conservative Forces F - Montgomery College
    The work done by a conservative force depends only on the beginning and ending positions of the object. Hence (as shown in the figure), the work done by the ...
  23. [23]
    8.2: Conservative and Non-Conservative Forces
    The work done by a conservative force is independent of the path; in other words, the work done by a conservative force is the same for any path connecting two ...
  24. [24]
    [PDF] 8. Conservative Forces and Potential Energy
    In particular, for any conservative force, we can define the change in potential energy of an object as minus the work done by this force. In this course ...
  25. [25]
    [PDF] Energy 2 - Duke Physics
    The change in one form of energy is accompanied by an opposite change in the other. This is the defining property of conservative forces. Conservation of ...
  26. [26]
    7.4 Conservative Forces and Potential Energy - College Physics 2e
    Jul 13, 2022 · Use the work-energy theorem to show how having only conservative forces implies conservation of mechanical energy. Potential Energy and ...
  27. [27]
    Potential energy and conservative forces (article) | Khan Academy
    If a resistive force like friction does the negative work, then the kinetic energy transferred from the object is dissipated as thermal energy and sound.
  28. [28]
    [PDF] Lecture D8 - Conservative Forces and Potential Energy
    It is clear that the potential satisfies dV = −F · dr (the minus sign is included for convenience). There are two main consequences that follow from the ...
  29. [29]
    14 Work and Potential Energy (conclusion) - Feynman Lectures
    Physical work is expressed as ∫F⋅ds, called “the line integral of F dot ds,” which means that if the force, for instance, is in one direction and the object on ...
  30. [30]
    Potential energy - Richard Fitzpatrick
    Fourthly, potential energy is only defined to within an arbitrary additive constant. ... potential energy between two points represents the net energy ...
  31. [31]
    Gravitational potential energy - Richard Fitzpatrick
    Gravitational potential energy. ... Here, we have adopted the convenient normalization that the potential energy at infinity is zero.
  32. [32]
    chap19 - Galileo
    $$mgh$ , where h is the height above an arbitrary reference level; given that the choice of reference level is arbitrary, potential energy does not have an ...
  33. [33]
    13.3 Gravitational Potential Energy and Total Energy
    First, U → 0 as r → ∞ . The potential energy is zero when the two masses are infinitely far apart. Only the difference in U is important, so the choice of U = ...
  34. [34]
    The Law of Conservation of Energy and Examples - Albert.io
    Feb 28, 2023 · We've now introduced three values that all have the same unit – kinetic energy, potential energy, and work. Since they have the same unit of ...
  35. [35]
    Dimensional Formula of Potential Energy - BYJU'S
    Therefore, Potential Energy is dimensionally represented as [M1 L2 T-2]. ⇒ Check Other Dimensional Formulas: Dimensions of Young Modulus · Dimensions of ...
  36. [36]
    9.1: Potential Energy of a System - Physics LibreTexts
    Jun 15, 2023 · A simple system embodying both gravitational and elastic types of potential energy is a one-dimensional, vertical mass-spring system. This ...
  37. [37]
    19.1 Electric Potential Energy: Potential Difference - UCF Pressbooks
    On the submicroscopic scale, it is more convenient to define an energy unit called the electron volt (eV), which is the energy given to a fundamental charge ...
  38. [38]
    Gravitational Potential Energy | Physics - Lumen Learning
    The gravitational potential energy of an object near Earth's surface is due to its position in the mass-Earth system. Only differences in gravitational ...Missing: limitations | Show results with:limitations
  39. [39]
    standard acceleration of gravity - CODATA Value
    standard acceleration of gravity $g_{\rm n}$ ; Numerical value, 9.806 65 m s ; Standard uncertainty, (exact).
  40. [40]
    AST 101: Forces, Orbits, and Energy
    The potential energy available due to gravity is given by the expression. U = -G m M / d ,. where m and M are the masses of the 2 objects and d is the ...Missing: formula | Show results with:formula
  41. [41]
  42. [42]
    14. GRAVITY - Home Page of Frank LH Wolfs
    Gravitational Potential Energy. In chapter 8 we have discussed the relation between the force and the potential energy. Consider two particles of masses m1 and ...
  43. [43]
    Week 8: Potential Energy and Energy Conservation | Physics
    Lesson 25: Potential Energy Diagrams · 25.1 Force is the Derivative of Potential · 25.2 Stable and Unstable Equilibrium Points · 25.3 Reading Potential Energy ...
  44. [44]
    Gravitational Potential Energy - HyperPhysics
    Since the zero of gravitational potential energy can be chosen at any point (like the choice of the zero of a coordinate system), the potential energy at a ...
  45. [45]
    [PDF] The Gravitational Energy of a Black Hole - arXiv
    An exact energy expression for a physical black hole is derived by considering the escape of a photon from the black hole. The mass of the black hole within.Missing: convention | Show results with:convention
  46. [46]
    Sun, Moon, Oceans: The Potential of Ocean Tidal Energy - Stanford
    Dec 10, 2019 · Tidal potential energy involves harnessing the potential energy stored in the oceans due to the height difference of high and low tides, whilst ...
  47. [47]
    Relativity and the Global Positioning System - Physics Today
    May 1, 2002 · If the GPS orbits were perfectly circular, the corrections would include just a few constant contributions: for the gravitational potential ...Gps Receivers · Rotating And Inertial Local... · The Topex Experiment
  48. [48]
    Elastic Potential Energy - HyperPhysics
    According to Hooke's law, the force required to stretch the spring will be directly proportional to the amount of stretch. Since the force has the form. F ...
  49. [49]
    16.1 Hooke's Law: Stress and Strain Revisited - UCF Pressbooks
    The potential energy stored in a spring is PE el = 1 2 k x 2 . Here, we generalize the idea to elastic potential energy for a deformation of any system that can ...Missing: formula | Show results with:formula
  50. [50]
    16.5 Energy and the Simple Harmonic Oscillator | Texas Gateway
    As the object starts to move, the elastic potential energy is converted to kinetic energy, becoming entirely kinetic energy at the equilibrium position. It is ...
  51. [51]
    The Simple Pendulum - Ximera - The Ohio State University
    When the angle is small, perhaps less than , then . This is referred to as the small angle approximation. This greatly simplifies the differential equation:.
  52. [52]
    [PDF] Strain Energy in Linear Elastic Solids - Duke People
    This last integral reduces to a constant that depends only upon the shape of the cross-section. This constant is given the variable name α. α = A. I2 z. ZZ. A.
  53. [53]
    [PDF] 16. Energy methods
    Write down the equilibrium form of the work energy equation for the system. For equilibrium, we know that P = ke for all deformations e. Therefore, the work ...Missing: numerical compute field
  54. [54]
    Young's Modulus as a Spring Constant - Stanford CCRMA
    Young's modulus is the Hooke's-law spring constant for the spring made from a specifically cut section of the solid material, cut to length 1 and cross- ...
  55. [55]
    [PDF] Part VII: Lattice vibrations – phonons 1 The simple harmonic oscillator
    Hamiltonian contains the kinetic energy (first term) and the potential energy (second term), which here is the elastic energy stored in the deformed spring.<|control11|><|separator|>
  56. [56]
    (PDF) The Ogden model of rubber mechanics: 50 years of impact on ...
    We place the Ogden model of rubber elasticity, published in Proceedings of the Royal Society 50 years ago, in the wider context of the theory of nonlinear ...
  57. [57]
  58. [58]
    The Feynman Lectures on Physics Vol. II Ch. 8: Electrostatic Energy
    We also know, from the principle of superposition, that if we have many charges present, the total force on any charge is the sum of the forces from the others.
  59. [59]
    8.2 Ionic Bonding
    The electrostatic attraction energy between ions of opposite charge is directly proportional to the charge on each ion (Q 1 and Q 2 in Equation 8.1). Thus, more ...
  60. [60]
    Electrostatic Energy - Richard Fitzpatrick
    This expression implies that the potential energy of a continuous charge distribution is stored in the electric field generated by the distribution ...
  61. [61]
    [PDF] Electric potential invariants and ions-in-molecules effective ...
    Dec 15, 2016 · In perturbation theory, the energy of the molecular ion modified by the external potential is a sum of terms: the energy of the static ...
  62. [62]
    Energy density within a dielectric medium - Richard Fitzpatrick
    The electrostatic energy density inside a dielectric medium is given by \begin{displaymath} U = \frac{1}{2} {\bf E}\!\cdot\!{\bf D}. \end{displaymath}
  63. [63]
    The Feynman Lectures on Physics Vol. II Ch. 10: Dielectrics - Caltech
    10–4The electrostatic equations with dielectrics. Now let's combine the above result with our theory of electrostatics. The fundamental equation is ∇⋅E=ρϵ0.
  64. [64]
    [PDF] 08. Electric potential and potential energy - DigitalCommons@URI
    Sep 25, 2020 · By construction, the potential energy has the value zero at this position. We are free to choose the reference position. Some choices are more ...
  65. [65]
    [PDF] Chapter 4 The Electric Potential
    Electric potential (V) is a function derived by dividing potential energy by charge, and its units are J/C (joules per coulomb).
  66. [66]
    Magnetic Potential Energy - HyperPhysics
    Magnetic potential energy depends on a magnetic dipole's orientation in a field, and is lowest when aligned. It relates to magnetic torque on a current loop.
  67. [67]
    The magnetic dipole moment - Physics
    For a coil with N turns, m = NIA. The torque is then given by: t = m ´ B. A loop, or coil, in a uniform magnetic field has a potential energy given by: U = -m · ...
  68. [68]
    [PDF] Magnetic Dipoles Magnetic Field of Current Loop i - MRI Questions
    Oct 24, 2006 · The potential energy on one dipole from the magnetic field from the other is: 1. 2. 1. 2 z z. U. B μ. = − ⋅. = − μ B. (choosing the z-axis for ...
  69. [69]
    [PDF] 32-1 gauss' law for magnetic fields - UF Physics
    The simplest magnetic structure that can exist and thus be enclosed by a Gaussian surface is a dipole, which consists of both a source and a sink for the field ...Missing: NMR | Show results with:NMR
  70. [70]
    12.7 Magnetism in Matter – University Physics Volume 2
    Paramagnetic materials have partial alignment of their magnetic dipoles with an applied magnetic field. This is a positive magnetic susceptibility. Only a ...
  71. [71]
    Nuclear Magnetic Resonance - HyperPhysics
    The magnetic potential energy difference is hυ = 2μB. The short table of Larmor frequencies below is from Hobbie, Ch 17 and Becker. An extensive list including ...
  72. [72]
    [PDF] Electromagnetic Fields and Energy - Chapter 9: Magnetization
    May 9, 2022 · The sources of magnetic field in matter are the (more or less) aligned magnetic dipoles of individual electrons or currents caused by ...
  73. [73]
    Magnetic energy - Richard Fitzpatrick
    The energy stored in an inductor is actually stored in the surrounding magnetic field. Let us now obtain an explicit formula for the energy stored in a ...
  74. [74]
    [PDF] Inductance and Magnetic Energy
    This coefficient L is called the self-inductance of the coil, which is often shortened to the coil's inductance or inductivity. Now let the current through the ...
  75. [75]
    17 The Laws of Induction - Feynman Lectures - Caltech
    But so long as the currents are changed slowly, the magnetic field will at each instant be nearly the same as the magnetic field of a steady current. We will ...
  76. [76]
    The Basics of Covalent Bonding in Terms of Energy and Dynamics
    We summarize the mechanistic bonding models and the debate over the last 100 years, with specific applications to the simplest molecules: H 2 + and H 2.
  77. [77]
    Types of Molecular Bonds – University Physics Volume 3
    An ionic bond forms when an electron transfers from one atom to another. A covalent bond occurs when two or more atoms share electrons. A van der Waals bond ...
  78. [78]
    Chemical Bonding in the H 2 Molecule
    For comparision, the experimental H-H bond length is 0.74 Å and the experimental bond dissociation energy is 436 kJ/mole or 4.52 eV. H-H Chemical Bond. As ...Missing: mol | Show results with:mol
  79. [79]
    [PDF] The Potential Energy Surface (PES)
    • But, the PES should not depend on the absolute location of the atoms, only on their location relative to one another (i.e., the molecular geometry). • So ...
  80. [80]
    [PDF] The Born-Oppenheimer Approximation
    equation for the motion of the nuclei on a given Born-Oppenheimer potential energy surface: [ ˆ. TN + T00 kk + Ukk] χk(R) = Eχk(R). (25). This equation clearly ...
  81. [81]
    [PDF] Density Functional Theory - Department of Chemistry
    Hybrid DFT was a breakthrough. Bond. Barrier energies heights. Hartree-Fock theory ... and algorithms to perform a HF and DFT calculation, the cost of DFT is ca.
  82. [82]
    Forms of energy - U.S. Energy Information Administration (EIA)
    Chemical energy is energy stored in the bonds of atoms and molecules. Batteries, biomass, petroleum, natural gas, and coal are examples of chemical energy.
  83. [83]
    Chemical Energy Storage | PNNL
    Chemical energy storage involves storing energy in chemical bonds, like hydrogen, which can be stored as a gas or liquid, or in materials, and can be converted ...
  84. [84]
    [PDF] Exploring Potential Energy Surfaces for Chemical Reactions
    Potential energy surfaces are central to studying molecular structures and reactivities. Methods include geometry optimization, transition state searching, and ...
  85. [85]
    Gibbs Free Energy
    The Gibbs free energy of a system at any moment in time is defined as the enthalpy of the system minus the product of the temperature times the entropy of the ...
  86. [86]
    [PDF] Combustion Chemistry - Princeton University
    Jun 24, 2018 · Combustion involves the oxidation of a fuel, ideally leading, for an organic fuel such as octane or ethanol, to the formation of carbon dioxide ...
  87. [87]
    Photosynthesis - The Cell - NCBI Bookshelf - NIH
    Photosynthesis takes place in two distinct stages. In the light reactions, energy from sunlight drives the synthesis of ATP and NADPH, coupled to the formation ...
  88. [88]
    Effect of catalysts - Chemistry 302
    The effect of a catalyst is that it lowers the activation energy for a reaction. Generally, this happens because the catalyst changes the way the reaction ...
  89. [89]
    Calorimetry - Chemistry 301
    Calorimetry is the measurement of heat, typically done at constant pressure or volume, using water temperature changes to quantify heat.
  90. [90]
    [PDF] Hideki Yukawa - Nobel Lecture
    The meson theory started from the extension of the concept of the field of force so as to include the nuclear forces in addition to the gravitational and.
  91. [91]
    [PDF] DAMTP - 3 The Strong Force
    Both the strong and weak nuclear forces share a common property with electromag- netism: the force is carried by a field of spin 1. In the case of ...
  92. [92]
    On the Neutron-Proton Scattering Cross Section | Phys. Rev.
    The Yukawa well is in better agreement with experiment than is the square well. One of the constants entering into the determination of the potential is the ...
  93. [93]
    Don't get too close | RIKEN
    Oct 12, 2007 · The repulsive core explains scattering experiments, the stability of nuclei, and even supernova explosions. It is probably caused by the ...
  94. [94]
    Woods-Saxon-type of mean-field potentials with effective mass ...
    Dec 30, 2020 · Those ready-for-use potentials are advertised as an alternative to other existing phenomenological mean-field potentials.
  95. [95]
    A.42 Nuclear forces - FAMU-FSU College of Engineering
    Because of the exponential in the Yukawa potential, the nuclear force is very short range. It is largely gone beyond distances of a couple of fm.
  96. [96]
    Yukawa potential approach to the nuclear binding energy formula
    Apr 1, 1990 · Yukawa's meson theory of the nucleon–nucleon interaction is used to show that the potential energy contains a volume term that is linear in the ...<|control11|><|separator|>
  97. [97]
    Energy and Nuclear Change - Chemistry 302
    The potential energy of a nucleus is simply the energy of the nucleus compared to the energy of the sub-atomic parts of the nucleus broken apart. For example, a ...
  98. [98]
    21.2: Nuclear Binding Energies - Physics LibreTexts
    Nov 6, 2024 · When nucleons are close-packed, the binding energy per nucleon due to the strong force is simply the number of nearest neighbors for each ...
  99. [99]
    Nuclear Binding Energy - HyperPhysics
    Nuclear binding energy is the difference between the mass of a nucleus and the sum of its protons and neutrons, holding the nucleus together.Missing: potential | Show results with:potential
  100. [100]
    The Isotopes of Hydrogen
    Aug 9, 2000 · The binding energy per nucleon continues to grow as protons and neutrons are added to construct more massive nuclei until a maximum of about 8 ...
  101. [101]
    (S-8A-2) Nuclear Binding Energy - PWG Home - NASA
    Feb 11, 2009 · The curve of binding energy (drawing) plots binding energy per nucleon against atomic mass. It has its main peak at iron and then slowly ...