Fact-checked by Grok 2 weeks ago

Potential gradient

In physics, chemistry, and , a potential gradient is the local rate of change of a with respect to , resulting in a that indicates the direction and magnitude of the steepest increase in the potential. This concept is particularly fundamental in , where it refers to the of the V, defined as the per unit positive test charge, measured in volts (). The points toward the region of increasing potential, with its magnitude quantifying the steepness of that spatial variation within an . The \mathbf{E} is related to the by \mathbf{E} = -\nabla V, meaning the field points in the direction of the most rapid decrease in potential, with magnitude equal to that rate of decrease. In Cartesian coordinates, this is expressed as E_x = -\frac{\partial V}{\partial x}, E_y = -\frac{\partial V}{\partial y}, E_z = -\frac{\partial V}{\partial z}. The units are volts per meter (V/m), equivalent to newtons per (N/C), describing the force on a charge in the field. This relation highlights the conservative nature of electrostatic fields, where work to move a charge depends only on potential difference, independent of path. Potential gradients apply beyond electrostatics to other conservative fields, such as gravitational and fluid potentials. In , the vertical electric potential gradient under fair weather conditions is approximately 100 V/m near the Earth's surface, arising from the global electric circuit with the surface negatively charged relative to the , yielding a total potential difference of about 400,000 V to the upper atmosphere. Such measurements aid in studying thunderstorms and , as field disruptions indicate approaching storms. In , potential gradients in monitor by revealing electrochemical activity and rates.

Definition

One-dimensional case

In one dimension, the \phi(x) is a that assigns a value to each x along a line, representing the per unit mass or charge at that point in a conservative . This potential is defined up to an additive constant, as only differences in \phi contribute to physical effects like work done by the . The potential in one dimension is the ordinary \frac{d\phi}{dx}, which quantifies the rate of change of the with respect to along the line. This indicates how steeply the potential varies at a given point x. Physically, the potential \frac{d\phi}{dx} points in the direction of steepest ascent of the potential (positive if increasing to the right, negative if to the left), while its measures the intensity of the associated per unit mass or charge. In conservative fields, this relates directly to the field's strength, motion toward regions of lower potential. A simple example is the linear potential \phi(x) = kx, where k is a constant; here, the gradient is \frac{d\phi}{dx} = k, corresponding to a constant force throughout the domain. For conservative fields in one dimension, the force F acting on a unit mass or charge is derived as the negative of the potential gradient: F = -\frac{d\phi}{dx}. This relation ensures that the work done by the force equals the negative change in potential, \int F \, dx = -\Delta \phi, conserving mechanical energy along the path.

Multidimensional case

In the multidimensional case, the potential gradient generalizes to a derived from a function \phi(\mathbf{r}), where \mathbf{r} is the position vector in n-dimensional . This \phi varies spatially, and its \nabla \phi captures the directional rate of change across multiple coordinates. In three-dimensional Cartesian coordinates, the gradient is expressed as the vector \nabla \phi = \left( \frac{\partial \phi}{\partial x}, \frac{\partial \phi}{\partial y}, \frac{\partial \phi}{\partial z} \right), where each component represents the partial derivative with respect to the respective coordinate. The gradient vector \nabla \phi points in the direction of the maximum rate of increase of \phi and is orthogonal to the equipotential surfaces (level sets where \phi is constant). Its magnitude is given by |\nabla \phi| = \sqrt{ \left( \frac{\partial \phi}{\partial x} \right)^2 + \left( \frac{\partial \phi}{\partial y} \right)^2 + \left( \frac{\partial \phi}{\partial z} \right)^2 }, which quantifies the steepness of the potential's variation at a point. The expression for the gradient transforms under different coordinate systems to account for the geometry of the space. In spherical coordinates (r, \theta, \phi), where r is the radial distance, \theta is the polar angle, and \phi is the azimuthal angle, the gradient takes the form \nabla \phi = \frac{\partial \phi}{\partial r} \hat{e}_r + \frac{1}{r} \frac{\partial \phi}{\partial \theta} \hat{e}_\theta + \frac{1}{r \sin \theta} \frac{\partial \phi}{\partial \phi} \hat{e}_\phi, with \hat{e}_r, \hat{e}_\theta, \hat{e}_\phi as the corresponding unit vectors. A representative example is the gravitational potential \phi(\mathbf{r}) = -\frac{GM}{r} in three dimensions, where G is the gravitational constant and M is a central mass. Its gradient is \nabla \phi = \frac{GM}{r^2} \hat{e}_r, directed radially outward with magnitude decreasing as the inverse square of the distance. The one-dimensional case arises as a special instance along a single axis, such as the radial direction here.

Applications in physics

Gravitational field

In Newtonian gravity, the gravitational field \mathbf{g} at a point in space is defined as the negative gradient of the gravitational potential \phi_\text{grav}, such that \mathbf{g} = -\nabla \phi_\text{grav}. This relationship arises because the gravitational force on a test mass m is conservative, expressible as \mathbf{F} = m \mathbf{g} = -m \nabla \phi_\text{grav}, ensuring the work done by gravity is path-independent. For a point mass M at the origin, the gravitational potential is \phi_\text{grav} = -\frac{GM}{r}, where G is the gravitational constant and r is the distance from the mass. The corresponding gravitational field is then \mathbf{g} = -\nabla \phi_\text{grav} = -\frac{GM}{r^2} \hat{e}_r, directed radially inward toward the mass. This concept originates from Isaac Newton's law of universal gravitation, published in 1687, which describes the force between masses but was later reformulated in terms of potentials to simplify calculations in celestial mechanics. Pierre-Simon Laplace formalized the use of gravitational potentials in the late 18th century, introducing them systematically in his work Mécanique Céleste (1799–1825) as part of potential theory to analyze planetary perturbations. For distributed mass densities, the gravitational potential satisfies Poisson's equation, \nabla^2 \phi_\text{grav} = 4\pi G \rho, where \rho is the mass density; in regions without mass (\rho = 0), it reduces to Laplace's equation \nabla^2 \phi_\text{grav} = 0. This differential form allows computation of \phi_\text{grav} and thus \mathbf{g} for complex mass distributions, such as stars or planets. A key example is a uniform solid sphere of radius R and total mass M. Outside the sphere (r > R), the field is identical to that of a point mass at the center: \mathbf{g} = -\frac{GM}{r^2} \hat{e}_r. Inside the sphere (r < R), the field is linear with distance: \mathbf{g} = -\frac{GM r}{R^3} \hat{e}_r, arising from the enclosed mass fraction within radius r. This contrast highlights how the potential gradient accounts for the superposition of contributions from all mass elements, vanishing at the center where symmetry balances forces.

Electric field

In electrostatics, the \mathbf{E} at any point in space is defined as the negative of the V, expressed mathematically as \mathbf{E} = -\nabla V. This relationship indicates that the points in the of the steepest decrease in potential and has a equal to the of change of the potential in that . The potential V is a scalar quantity measured in volts, representing the work done per unit charge to bring a test charge from a reference point (often ) to that location. For a single point charge q located at the origin, the electric potential at a distance r from the charge is given by V = \frac{q}{4\pi\epsilon_0 r}, where \epsilon_0 is the . Taking the negative gradient of this potential yields the corresponding \mathbf{E} = \frac{q}{4\pi\epsilon_0 r^2} \hat{r}, directed radially outward for a positive charge and following the . This derivation confirms the consistency between potential and field descriptions for isolated charges. The connection between the electric field and charge distribution arises from Gauss's law in differential form, \nabla \cdot \mathbf{E} = \frac{\rho}{\epsilon_0}, where \rho is the . Substituting \mathbf{E} = -\nabla V into this equation produces , \nabla^2 V = -\frac{\rho}{\epsilon_0}, which governs the potential in regions with nonzero . In charge-free regions, \rho = 0, simplifying to \nabla^2 V = 0. On the surface of a in electrostatic , the V is constant, implying that the tangential component of \mathbf{E} vanishes and is to the surface. This boundary condition arises because any potential difference along the surface would induce currents until is reached, with excess charge residing on the outer surface. A practical example is the parallel-plate capacitor, consisting of two oppositely charged conducting plates separated by distance d. Between the plates, the electric field is uniform and directed from the positive to the negative plate, with magnitude E = -\frac{\Delta V}{d}, where \Delta V is the potential difference across the plates. This uniform field approximation holds when the plate separation is much smaller than their dimensions, enabling straightforward calculations of capacitance and energy storage.

Pressure and velocity potentials in fluids

In , the concept of potential gradient manifests in through the relationship between and . For a at rest under , the p satisfies the \nabla p = -\rho \nabla \Phi, where \rho is the and \Phi is the , typically \Phi = gz near Earth's surface with g as the and z the vertical coordinate. This gradient implies that the force per unit volume on the is -\rho \nabla p = \rho \nabla \Phi, balancing the gravitational , and is often simplified to \nabla p = -\rho g in the vertical for constant fluids. For dynamic flows, potential gradients appear prominently in the for irrotational, incompressible fluids. In such cases, the velocity field \mathbf{v} can be expressed as the of a scalar velocity potential \phi, so \mathbf{v} = \nabla \phi, which inherently satisfies the irrotational condition \nabla \times \mathbf{v} = 0. This representation simplifies the governing s, leading to Bernoulli's equation along streamlines for steady, : \frac{1}{2} v^2 + \frac{p}{\rho} + gz = \ constant, where the terms connect per unit mass, , and . Substituting v = |\nabla \phi|, the equation highlights how gradients of \phi and the \Phi = gz interplay with variations. In source-free regions of steady, incompressible, irrotational flow, the velocity potential \phi obeys \nabla^2 \phi = 0, derived from the \nabla \cdot \mathbf{v} = 0 and \mathbf{v} = \nabla \phi. Solutions to this describe harmonic fields, analogous to , and enable analytical prediction of flow patterns in ideal fluids. A classic example is the potential flow around a in a , modeled by superposing a flow potential \phi_u = -U z (with U as the far-field speed) and a potential \phi_d = \frac{\mu \cos \theta}{r^2} centered at the , where \mu = U a^3 with a the , r the radial distance, and \theta the polar angle. The total potential \phi = -U r \cos \theta + \frac{U a^3 \cos \theta}{r^2} yields the via \mathbf{v} = \nabla \phi, which is tangent to the surface (no penetration) and approaches the far away, illustrating stagnation points at the poles where \nabla \phi = 0. The foundational ideas for these potential-based descriptions trace to Leonhard Euler's 18th-century work on ideal (inviscid, incompressible) fluids, where he derived the and introduced the for irrotational flows in his 1757 treatise Principia motus fluidorum. Euler's formulations laid the groundwork for modern theory, emphasizing conservative forces derivable from scalar potentials.

Applications in other sciences

Chemical potential gradients

In chemical systems, the chemical potential \mu_i of species i represents the partial molar , quantifying the energy change associated with adding one of that species to the system at constant temperature, pressure, and composition. The gradient of the chemical potential, \nabla \mu_i, serves as the fundamental driving force for diffusive , as species move from regions of higher to lower chemical potential to minimize . In , the diffusive flux J_i of species i is proportional to this gradient, given by J_i = -\frac{c_i D_i}{RT} \nabla \mu_i, where c_i is the concentration, D_i the diffusion coefficient, R the , and T the temperature; this formulation, known as the Nernst-Planck equation in terms of chemical potential, generalizes Fick's laws by accounting for thermodynamic driving forces beyond mere concentration differences. For charged species in electrochemical systems, the relevant potential is the electrochemical potential \tilde{\mu}_i = \mu_i + z_i F V, where z_i is the , F is the , and V is the ; its gradient \nabla \tilde{\mu}_i incorporates both chemical (concentration and activity) and electrical contributions, driving ion fluxes under combined diffusive and migratory forces. The flux then becomes J_i = -\frac{c_i D_i}{RT} \nabla \tilde{\mu}_i, enabling the description of transport in solutions where electric fields arise from charge imbalances. In ideal dilute solutions without electric fields, the chemical potential approximates \mu_i \approx \mu_{0,i} + RT \ln c_i, where \mu_{0,i} is the standard chemical potential; substituting this into the flux expression yields \nabla \mu_i \approx \frac{RT}{c_i} \nabla c_i, simplifying to Fick's first law J_i = -D_i \nabla c_i, which highlights how concentration gradients approximate chemical potential gradients under these conditions. A key example of chemical potential gradients in action is the Donnan equilibrium, where a semipermeable membrane separates a solution containing mobile ions from one with fixed charged macromolecules, leading to unequal ion distributions; here, the chemical potential gradients for permeant ions are balanced by an induced electric potential difference (Donnan potential), achieving equilibrium when electrochemical potentials are equal across the membrane. In practical applications, such as lithium-ion batteries, spatial gradients in for ions across electrodes and electrolytes drive intercalation and deintercalation processes during charge-discharge cycles, generating the cell voltage as the difference in electrochemical potentials between and . These gradients limit battery performance at high rates, as steep profiles increase overpotentials and reduce efficiency.

Biological membrane potentials

Biological membrane potentials arise from the uneven distribution of ions across the of membranes, creating a voltage that is essential for cellular function. In neurons and other excitable s, the resting is typically around -70 mV, with the interior of the negative relative to the exterior. This is primarily established and maintained by the Na⁺/K⁺ ATPase pump, which actively transports three sodium ions out of the and two ions in, counteracting passive ion leaks and sustaining the necessary concentration differences. Action potentials represent dynamic alterations in this potential gradient, enabling rapid . Upon reaching a , voltage-gated sodium channels open, allowing a massive influx of Na⁺ ions that depolarizes the from -70 mV toward +30 mV in milliseconds. This rapid change reverses the potential gradient temporarily, and the associated across the thin (~5-10 nm) is given by \mathbf{E} = -\nabla V, where V is the , driving ion movements that propagate the impulse along the . Subsequent potassium efflux through voltage-gated channels repolarizes the , restoring the resting state. The magnitude of these ion-specific potential gradients is predicted by the Nernst equation, which calculates the equilibrium voltage for a given based on its concentration ratio across the : V_{\text{ion}} = \frac{RT}{zF} \ln \left( \frac{[\text{ion}]_{\text{out}}}{[\text{ion}]_{\text{in}}} \right) Here, R is the , T is , z is the ion's , and F is Faraday's constant; for , this yields approximately -90 mV, influencing the overall . These gradients are intrinsically tied to differences for ions, powering selective transport. Potential gradients underpin neuronal signaling by providing the electrochemical driving force for , which travel as impulses along axons at speeds up to 100 m/s, and for synaptic transmission, where triggers release into the synaptic cleft. At chemical synapses, the arriving alters the postsynaptic , either exciting or inhibiting the next based on the gradient's direction and strength. The Hodgkin-Huxley model, developed in 1952 from experiments on squid giant axons, quantitatively describes these processes through differential equations governing voltage-gated Na⁺ and K⁺ conductances, predicting how channel kinetics generate and sustain the potential dynamics during . This framework revolutionized understanding of excitable membranes and remains foundational for .

Mathematical aspects

Non-uniqueness of scalar potentials

In conservative fields, the scalar potential \phi is defined only up to an arbitrary additive constant C, meaning that \phi and \phi + C produce identical gradients \nabla \phi = \nabla (\phi + C). This non-uniqueness arises because the physical quantities of interest, such as forces and field strengths, depend solely on the spatial variation of the potential rather than its absolute value. A key consequence is the path independence of line integrals in such fields. For a conservative vector field \mathbf{F} = \nabla \phi, the line integral \int_A^B \mathbf{F} \cdot d\mathbf{r} = \phi(B) - \phi(A), which depends only on the endpoints and not on the specific path taken, as guaranteed by the fundamental theorem for line integrals. This property holds provided the domain is simply connected and \mathbf{F} is the gradient of a scalar potential, ensuring that closed-path integrals vanish. This ambiguity in the mirrors gauge freedom in physics, where transformations that leave observable fields unchanged are permissible, though for static scalar cases, the freedom is limited to adding a constant unlike the more general time-dependent gauges in . Consequently, physical forces \mathbf{F} = -q \nabla \phi (for a charge q in an , or analogously \mathbf{F} = -m \nabla \phi in ) remain invariant, as only potential differences or the itself determine the field's effects. For instance, in the , the choice of zero point for the potential—whether at or Earth's surface—does not alter orbital dynamics, since the depend on \nabla \phi alone, preserving Keplerian orbits regardless of the additive .

Connection to potential theory

is a branch of that investigates the properties of harmonic functions, which satisfy ∇²φ = 0, and their gradients, providing a foundational framework for understanding potential gradients in physical contexts such as and gravitation. These gradients represent the directional derivatives of the scalar potential φ, yielding vector fields that model conservative forces, as the of the gradient vanishes (∇ × ∇φ = 0). The theory extends to inhomogeneous cases via ∇²φ = -f, where f denotes a source term, linking directly to the computation of potential gradients in the presence of distributed charges or masses. A key tool in for solving is the use of , which allow the potential to be expressed as an integral: φ(r) = ∫ G(r, r') f(r') dV', where G(r, r') is the Green's function satisfying ∇²G = δ(r - r') with appropriate boundary conditions. The of this potential then yields the field strength, enabling analytical solutions for simple geometries and serving as a basis for more complex derivations in potential gradient problems. This approach, originally developed by George Green in 1828, underpins much of modern by transforming differential equations into integral forms that are computationally tractable. Multipole expansions further connect potential gradients to by decomposing the potential φ into a series of terms, starting from the (proportional to 1/r), whose gradient produces the basic ; higher-order terms like and quadrupoles arise from gradients of these multipoles, capturing the directional variations in the for distant sources. In , for instance, the is given by the gradient of (p · r)/r³, illustrating how organizes the spatial decay and angular dependence of gradients. These expansions, formalized in the , facilitate approximations for systems with multiple sources, essential for analyzing potential gradients in non-uniform distributions. The historical development of potential theory traces back to in the early , who applied it to through his 1835 work on the intensity of electrical forces, deriving the 1/r² law for gradients from the harmonic properties of the potential. Gauss's contributions were later extended to gravitational potentials by and others, establishing potential theory as a unified mathematical structure for both electromagnetic and gravitational gradients, influencing fields like . This framework evolved through the with rigorous treatments by mathematicians such as Oliver Kellogg, who emphasized boundary value problems for harmonic functions. In applications, potential theory supports numerical methods for computing complex potential gradients, such as the (FEM), which discretizes the domain to solve Laplace or equations approximately and then evaluates gradients via . FEM is particularly valuable for irregular boundaries where analytical Green's functions are unavailable, allowing efficient simulation of gradient fields in contexts like electromagnetic design; for example, it has been used to model potential gradients in high-voltage systems achieving high accuracy with refined meshes. These methods build on 's variational principles, minimizing energy functionals to ensure gradient accuracy.

References

  1. [1]
    Electric Potential and Electric Field - Richard Fitzpatrick
    The electric field is the negative gradient of electric potential. Electric potential is a scalar field, while the electric field is a vector field.Missing: physics | Show results with:physics
  2. [2]
    [PDF] Chapter 3 Electric Potential - MIT
    ... potential) and results in a vector quantity (electric field). Mathematically, we can think of E as the negative of the gradient of the electric potential V .
  3. [3]
    [PDF] Lecture 11: Potential Gradient and Capacitor
    A capacitor is a device with separated conductors that store charge and energy. Potential gradient points where potential increases fastest. Capacitance ...
  4. [4]
    9 Electricity in the Atmosphere - Feynman Lectures - Caltech
    Another thing that can be measured, in addition to the potential gradient, is the current in the atmosphere. The current density is small—about 10 ...Missing: definition | Show results with:definition
  5. [5]
    Potential Gradient - an overview | ScienceDirect Topics
    Potential gradient refers to the variation of electrical potential across a given area, particularly in the context of corrosion systems in reinforced concrete, ...
  6. [6]
    [PDF] Lecture D8 - Conservative Forces and Potential Energy
    When operating on a scalar function V (x, y, z), the result VV is a vector, called the gradient of V . The components of VV are the derivatives of V along each ...
  7. [7]
    [PDF] Lecture 8 Gradient fields/conservative forces - the waterloo
    A conservative force is the negative derivative of a potential energy function, and in higher dimensions, it can be defined as F = -∇V.
  8. [8]
    Calculus III - Vector Fields - Pauls Online Math Notes
    Nov 16, 2022 · In this section we introduce the concept of a vector field and give several examples of graphing them. We also revisit the gradient that we ...
  9. [9]
    4.1 Gradient, Divergence and Curl
    The gradient of a scalar-valued function f ( x , y , z ) is the vector field. grad grad f = ∇ ∇ f = ∂ f ∂ x ı ı ^ + ∂ f ∂ y ȷ ȷ ^ + ∂ f ∂ z k ^ · The divergence ...
  10. [10]
    [PDF] Lecture 5 Vector Operators: Grad, Div and Curl
    We introduce three field operators which reveal interesting collective field properties, viz. • the gradient of a scalar field,. • the divergence of a vector ...
  11. [11]
    Spherical Coordinates - Richard Fitzpatrick
    $$\displaystyle dV = r^{\,2}\,\sin, (C.63). According to Equations (C.13), (C.15), and (C.18), gradient, divergence, and curl in the spherical coordinate system ...
  12. [12]
    [PDF] Gravity 3 - Gravitational Potential and the Geoid
    In this vector form we can think of gravitational acceleration in directions ... Gravitational acceleration is the gradient of the potential: g = −. GM.
  13. [13]
    [PDF] Vector Calculus in Two Dimensions
    The scalar field u is often referred to as a potential function for its gradient vector field ∇u. For example, the gradient of the potential function u(x ...
  14. [14]
    [PDF] 1 Gravitational and Electrostatic Potential Energy
    For both point-source potentials, the force on an object m or q can be expressed as the negative gradient of the potential energy. ~. F = −∇ U(r). (7). 2 ...
  15. [15]
    Potential Energy - HyperPhysics
    The force on an object is the negative of the derivative of the potential function U. This means it is the negative of the slope of the potential energy curve.
  16. [16]
    Gravitational Potential
    From Equation (1.264), the gravitational potential due to a point object (or a spherically symmetric object) of mass $M$ situated at the origin is ...
  17. [17]
    5.8.1: Potential Near a Point Mass - Physics LibreTexts
    Mar 5, 2022 · The potential at a point is the work required to move unit mass from infinity to the point; ie, it is negative.
  18. [18]
    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 ...
  19. [19]
    [PDF] 5 The Poisson and Laplace Equations - DAMTP
    The plot of Φ is shown on the left of Figure 17; the plot of the gravitational field g = dΦ/dr is on the right, where we see a linear increase inside the planet ...
  20. [20]
    Gravitational Field Intensity - BYJU'S
    Inside the uniform solid sphere (r<R), E = -GMr/R3. On the surface of the uniform solid sphere (r= R), E = -GM/R2. Outside the uniform solid sphere (r>R), E ...Gravitational Field Intensity · due to Ring · due to Uniform Spherical Shell
  21. [21]
  22. [22]
    Concept: The negative gradient of the electric potential is the electric ...
    The negative gradient of the electric potential is the electric field. While the magnitude of the electric field is equal to the gradient of the electric ...
  23. [23]
    Electric Field from Voltage - HyperPhysics
    The component of electric field in any direction is the negative of rate of change of the potential in that direction.
  24. [24]
    Electric potential for different charge geometries - HyperPhysics
    Point Charge Potential. The electric potential (voltage) at any point in space produced by a point charge Q is given by the expression below.
  25. [25]
    19.3 Electrical Potential Due to a Point Charge - UCF Pressbooks
    Electric potential of a point charge is V = k Q / r . Electric potential is a scalar, and electric field is a vector. Addition of voltages ...
  26. [26]
    Example 5.4: Electric potential due to point charges
    The change in electric potential energy of the third charge as it moves from its initial position to infinity is the product of the third charge, $q_c$.
  27. [27]
    Relation of Electric Field to Charge Density - HyperPhysics
    The divergence of the electric field at a point in space is equal to the charge density divided by the permittivity of space.
  28. [28]
    LaPlace's and Poisson's Equations - HyperPhysics
    The potential is related to the charge density by Poisson's equation. In a charge-free region of space, this becomes LaPlace's equation.
  29. [29]
    Poisson's equation - Richard Fitzpatrick
    Let us now try to develop some techniques for solving Poisson's equation which allow us to solve real problems (which invariably involve conductors). next · up ...
  30. [30]
    Electric Potential due to Conductors
    Inside the electric field vanishes. That means the electric potential inside the conductor is constant.
  31. [31]
    [PDF] 11. Electric potential of conductors, electric dipole, and point
    Oct 5, 2020 · The electric potential inside a conductor is the same at all points. The potential of a charged shell is like a point charge outside and  ...
  32. [32]
    Parallel-Plate Capacitor - Physics
    In the region between the plates and away from the edges, the electric field, pointing from the positive plate to the negative plate, is uniform. For a ...
  33. [33]
    PHYS208 Evaluation of Parallel-Plate Capacitance
    Mar 9, 1998 · The plates are considered to be large enough in area and close enough in distance that the electric field is uniform well away from the ...
  34. [34]
    Fluid Statics & the Hydrostatic Equation – Introduction to Aerospace ...
    Pressure forces acting on an infinitesimally small, stationary fluid volume with pressure gradients in three directions.
  35. [35]
    [PDF] Potential Flow Theory - MIT
    Bernoulli Equation​​ When flow is irrotational it reduces nicely using the potential function in place of the velocity vector. The potential function can be ...
  36. [36]
    Irrotational Flow - Richard Fitzpatrick
    In other words, the velocity potential in an incompressible irrotational fluid satisfies Laplace's equation. According to Equation (4.84), if the flow ...
  37. [37]
    [PDF] CHAPTER 10 ELEMENTS OF POTENTIAL FLOW ( )# !2U
    The dipole added to a uniform flow generates the potential flow about a sphere in uniform flow. Page 15. 15 !Sphere = !Uniform Flow + !Dipole = U" ...
  38. [38]
    Flow Past a Spherical Obstacle - Richard Fitzpatrick
    Consider the steady flow pattern produced when an impenetrable rigid spherical obstacle is placed in a uniformly flowing, incompressible, inviscid fluid.
  39. [39]
    [PDF] FROM NEWTON'S MECHANICS TO EULER'S EQUATIONS
    In retrospect, Euler was right in judging that his “two equations” were the definitive basis of the hydrodynamics of perfect fluids. He reached them at the end ...
  40. [40]
    Euler's Ideal Fluids:
    Euler spent some time occasionally investigating the theory of ideal fluids, at a time when the physical properties of liquids such as viscosity and surface ...
  41. [41]
    [PDF] Lecture 3: Diffusion: Fick's first law
    In most cases, chemical potential increases with increasing concentration, so it is convenient to express diffusion in term of concentration.
  42. [42]
    [PDF] 10.626 Lecture Notes, Diffuse charge in electrolytes - DSpace@MIT
    The flux density for the Nernst-Planck Equation can be generally expressed as ... and the gradient of the chemical potential for a dilute solution, we can ...
  43. [43]
    [PDF] Total Driving Force for Ionic Transport: Nernst-Planck Flux Equation
    chemical potential and electrostatic potential. The sum of the two ... ❖ Total Driving Force for Ionic Transport: Nernst-Planck Flux Equation ...
  44. [44]
    Designing interfaces in energy materials applications with first ...
    Feb 15, 2019 · Subsequently, μ(H+) + μ(e−) can be shifted to the desired potential by applying the Nernst's equation ΔG = −zFV (see later Eq. (10)), where V is ...
  45. [45]
    [PDF] Nernst-Planck Theory vs Modified Fick's Law - OSTI.GOV
    Oct 4, 2016 · Here, the gradient of the electrochemical potential in Eq. 47 is ex- pressed by Eq. 35, with A in the form of Eq. 48. With the assumption of ...<|separator|>
  46. [46]
    [PDF] The theory of membrane equilibria - Electrochemistry Knowledge
    In 1911 the author in collaboration with Harris (2) noticed the occurrence of these peculiar ionic equilibria during the course of an investigation of the ...
  47. [47]
    Brief overview of electrochemical potential in lithium ion batteries
    The gradient concentration directly results in the chemical potential ( μ Li + ) variation, which is one of the driving forces. And the inhomogeneous ...
  48. [48]
    [PDF] Transport and Kinetic Phenomena Linked to Power Performance of ...
    batteries are available, lithium-ion (Li-ion) rechargeable batteries dominate the market ... Lithium ion diffusion is driven by the chemical potential gradient.
  49. [49]
    Physiology, Resting Potential - StatPearls - NCBI Bookshelf - NIH
    Since the plasma membrane at rest has a much greater permeability for K+, the resting membrane potential (-70 to -80 mV) is much closer to the equilibrium ...
  50. [50]
    Physiology, Sodium Potassium Pump - StatPearls - NCBI Bookshelf
    Mar 13, 2023 · It has an ongoing role in stabilizing the cell's resting membrane potential, regulating cell volume and signal transduction.[2] It plays a ...Introduction · Cellular Level · Function
  51. [51]
    Physiology, Action Potential - StatPearls - NCBI Bookshelf - NIH
    At each node, the membrane depolarizes above the threshold voltage, and the influx of sodium ions again initiates the action potential through Nav.
  52. [52]
    Membrane Electromechanics in Biology, with a Focus on Hearing
    The electric field across the hair cell membrane changes within microseconds of bending the bundle. The field affects electromechanical force generation in the ...
  53. [53]
    Membrane Potentials, Synaptic Responses, Neuronal Circuitry ...
    The Nernst equation is generally considered for ions across a membrane generating an electromotive force as commonly shown as: V= (RT/zF) ln([X]out/ [X]in) X = ...
  54. [54]
    Ion Channels and the Electrical Properties of Membranes - NCBI - NIH
    Neuronal signals are transmitted from cell to cell at specialized sites of contact known as synapses. The usual mechanism of transmission is indirect. The ...
  55. [55]
    [PDF] Physics 4183 Electricity and Magnetism II Electrostatics and Ohm's ...
    Dec 17, 2012 · potential is defined up to a constant; in other words, only potential differences matter not the ... We first calculate the gradient of the scalar ...
  56. [56]
    [PDF] Electromagnetism with Spacetime Algebra - Wooster Physics
    Feb 27, 2024 · The scalar nature of the electrostatic source ρ suggests a scalar potential ϕ in the ... The electric potential is unique up to a constant ...
  57. [57]
    Calculus III - Fundamental Theorem for Line Integrals
    Nov 16, 2022 · If →F F → is conservative then it has a potential function, f f , and so the line integral becomes ∫C→F⋅d→r=∫C∇f⋅d→r ∫ C F → ⋅ d r → = ∫ C ∇ f ⋅ ...
  58. [58]
    [PDF] 18.02SC Notes: Fundamental Theorem for Line Integrals
    F = vf is a conservative field. Example 1: If F is the electric field of an electric charge it is conservative. Example 2: The gravitational field of a mass ...
  59. [59]
    [PDF] Maxwell's Equations - Scalettar
    This freedom to add a constant potential is called gauge freedom and the different potentials one can obtain that lead to the same physical field are ...
  60. [60]
    Gravitational potential energy - Richard Fitzpatrick
    Since potential energy is undetermined to an arbitrary additive constant, we could just as well write. \ ... Next: Satellite orbits Up: Orbital motion Previous: ...