Fact-checked by Grok 2 weeks ago

Quasistatic approximation

The quasistatic approximation is a in physics used to simplify the analysis of systems undergoing slow changes by treating the system as being in a static at each instant, neglecting time derivatives in the governing equations when variations occur on timescales much longer than intrinsic relaxation times. In , this approximation treats time-varying electric and magnetic fields as if they were static at each instant, assuming changes in sources propagate instantaneously without significant or effects. This approach is valid for systems where the characteristic of variations, τ, greatly exceeds the electromagnetic wave transit time across the system's dimensions, L/c (with c the ), typically requiring L/c ≪ τ. The approximation finds applications across various fields, including for processes that maintain internal equilibrium, for quasi-steady flows, and . In practice, within electromagnetism, it manifests in two primary forms: the electroquasistatic (EQS) approximation, which neglects time-varying (∂B/∂t term in Faraday's ) while retaining , and the magnetoquasistatic (MQS) approximation, which neglects the (∂D/∂t term in Ampère's ) while incorporating magnetostatics. Under EQS, the fields are primarily determined by charge distributions using and the electrostatic potential, with error fields scaling as (L/cτ)^2 relative to the dominant fields. Similarly, MQS focuses on distributions via Ampère's , applying to scenarios like inductors or conductors with alternating s. These reductions bridge the gap between static field theory and the full , enabling tractable solutions for low-frequency phenomena. Key applications in electromagnetism include analyzing the skin effect, where alternating currents confine to a thin layer near a conductor's surface due to eddy currents, with penetration depth δ = √(2/μσω) (μ permeability, σ conductivity, ω angular frequency). For instance, at 60 Hz in copper, δ ≈ 8.5 mm, but it shrinks to microns at gigahertz frequencies, impacting high-frequency devices like transformers and antennas. The approximation also underpins models in plasma physics, neuromodulation, and atmospheric dynamics, where slow variations allow neglecting wave propagation. Its limitations arise at higher frequencies, where full electrodynamic treatment becomes necessary to account for radiated energy.

Introduction

Definition

The quasistatic approximation is a modeling technique in physics that assumes a process occurs infinitely slowly relative to the system's characteristic relaxation times, ensuring the system remains in thermodynamic or at every instant. This approach idealizes the dynamics such that the system can be treated as traversing a continuous path through states, with any perturbations decaying rapidly enough to maintain near-equilibrium conditions. Key characteristics of the quasistatic approximation include the neglect of transient effects, such as rapid fluctuations or non-equilibrium transients, allowing the system's evolution to be described using thermodynamic relations or static field equations at each step. Consequently, time-dependent terms in the governing equations, like certain partial derivatives with respect to time, are deemed negligible when the process timescale far exceeds the system's internal response time. The concept originated in the within , where and contemporaries developed it to analyze the efficiency of heat engines through idealized reversible cycles. and cyclic processes in the 1850s and 1860s formalized the idea of processes composed of steps, distinguishing them from irreversible ones. In general, the quasistatic approximation finds broad applicability across physics disciplines, simplifying complex dynamic equations by ignoring small time derivatives, thereby bridging static and fully time-dependent analyses in fields ranging from to .

Validity conditions

The quasistatic approximation relies fundamentally on a clear separation of timescales between the duration of the external driving the system, denoted as \tau_\text{process}, and the intrinsic relaxation timescale \tau_\text{relax} over which the system returns to following a . This , \tau_\text{process} \gg \tau_\text{relax}, ensures that the system can maintain thermodynamic or at every stage, allowing state variables such as , , or fields to be well-defined throughout the evolution. Without this separation, deviations from accumulate, rendering the static-like description invalid. For a characterized by a typical size L and a relevant speed v—such as the in or the in electromagnetic contexts—the relaxation timescale is approximately \tau_\text{relax} \approx L / v. The holds when the rate of change imposed by the process is slow relative to the inverse of this timescale, i.e., when variations occur on times much longer than L / v, preventing wave-like effects from dominating. This establishes the boundary of applicability across diverse physical domains, where v sets the for how quickly disturbances equilibrate within the . The errors arising from the quasistatic approximation typically scale as \mathcal{O}\left((\tau_\text{relax} / \tau_\text{process})^2\right) in contexts like , which must be much less than unity to ensure quantitative accuracy; for instance, deviations in predicted fields or state variables remain small only if this ratio is sufficiently suppressed. This perturbative error bound underscores the approximation's utility for slow evolutions but highlights its sensitivity to the degree of timescale separation. A key limitation occurs near critical points or during phase transitions, where \tau_\text{relax} diverges to critical slowing down, eliminating the necessary timescale separation and causing the to exhibit non-equilibrium even for nominally slow processes. In such regimes, the quasistatic description fails, as the cannot adapt instantaneously to changes, leading to or other irreversible effects.

Thermodynamic applications

Quasistatic processes

In thermodynamics, a quasistatic process refers to an idealized transformation of a thermodynamic system, such as a gas, where the system maintains internal equilibrium at every stage, with state variables like pressure P, volume V, and temperature T changing infinitesimally slowly. This ensures that the process occurs on a timescale much longer than the system's relaxation time, allowing equilibrium thermodynamics to describe the system continuously without significant deviations from balance. On a pressure-volume (PV) diagram or temperature-entropy (TS) diagram, a quasistatic process traces a smooth, continuous path through the state space, differing from irreversible processes that feature abrupt jumps between non-equilibrium states. For an ideal gas undergoing such a process, the work done by the system is calculated as W = \int P \, dV, integrated along this equilibrium path, while the heat transfer Q is determined via the first law of thermodynamics: \Delta U = Q - W, where U is the change in internal energy. A classic example is the slow compression of an confined in a frictionless -cylinder device, where the piston advances at an infinitesimally slow rate, permitting the gas to remain in and exchange with its surroundings to adjust gradually during the volume reduction. This setup highlights the role of controlled in maintaining the quasistatic nature, ensuring the process can be analyzed using equilibrium properties throughout. Quasistatic processes form the basis for understanding reversible thermodynamic transformations, as explored in related contexts.

Relation to reversible processes

A quasistatic process is reversible if it occurs without dissipative effects, such as or , enabling both the system and its surroundings to be restored to their initial states without net change. This equivalence holds because the absence of ensures infinitesimal changes maintain , allowing the process to proceed equally well in reverse. In quasistatic reversible processes, the total entropy change of the , \Delta S_\text{total}, equals zero, representing the boundary case between reversible and irreversible behaviors. For irreversible processes, \Delta S_\text{total} > 0, indicating due to , whereas quasistatic conditions minimize but do not eliminate this in real scenarios. The exemplifies a quasistatic reversible , consisting of two isothermal and two adiabatic steps, all executed infinitely slowly to maintain . This configuration achieves the maximum theoretical for a operating between hot and cold reservoirs, given by \eta = 1 - \frac{T_\text{cold}}{T_\text{hot}}, where temperatures are in . In practice, real thermodynamic approximate quasistatic behavior only when conducted sufficiently slowly, yet they invariably include some irreversibility from dissipative mechanisms like . Thus, while quasistatic approximations model reversibility for analysis, actual implementations fall short, leading to reduced and generation.

Fluid dynamics applications

Quasi-steady flow approximation

The quasi-steady flow approximation in fluid dynamics applies to situations where the temporal variations in the flow field occur slowly compared to the convective timescale, allowing the neglect of unsteady terms in the Navier-Stokes equations. Specifically, the core assumption is that the partial derivative with respect to time, ∂u/∂t, can be omitted when the Strouhal number St = f L / U ≪ 1, where f is the characteristic frequency, L is the length scale, and U is the flow velocity; this indicates that the time for flow changes is much longer than the time for convection across the domain L/U. Under this approximation, the flow responds nearly instantaneously to changes in boundary conditions or forcing, treating the problem as locally steady at each instant. The governing equations simplify accordingly. For , the unsteady Navier-Stokes momentum equation \rho \left( \frac{\partial \mathbf{u}}{\partial t} + \mathbf{u} \cdot \nabla \mathbf{u} \right) = -\nabla p + \mu \nabla^2 \mathbf{u} + \mathbf{f} reduces to the steady form by ignoring the ∂u/∂t term: \rho (\mathbf{u} \cdot \nabla \mathbf{u}) \approx -\nabla p + \mu \nabla^2 \mathbf{u} + \mathbf{f}, where ρ is , p is , μ is dynamic , and f represents body forces; this yields the steady Euler equations for inviscid cases (μ=0) or steady Navier-Stokes for viscous flows. The remains unchanged as ∇ · u = 0 for incompressibility. This approximation holds for flows with low-frequency oscillations or slowly varying conditions, where the timescale separation ensures unsteady effects are negligible. It is particularly valid in flows, where the quasi-steady balance equates gradients from tidal head differences to frictional forces, neglecting inertial due to the long tidal period relative to flow transit times. Similarly, in creeping s at low Reynolds numbers (Re ≪ 1), such as the slow motion of small particles in viscous fluids, the quasi-steady applies because viscous equilibrates the flow much faster than boundary changes, allowing treatment of unsteady motions as a sequence of steady Stokes problems. A representative application is for thin-film flows, where the quasi-steady approximation balances pressure gradients with viscous shear forces across the thin gap, neglecting both inertial and unsteady terms to derive the for pressure distribution. In this regime, the flow in narrow channels or bearing films adjusts rapidly to surface motions, enabling efficient computation of load-carrying capacity without resolving full transient dynamics.

Applications in low-speed flows

In low-speed flows, the quasistatic approximation, often termed the quasi-steady approximation, is particularly relevant in the incompressible limit where the \rho remains constant and the \mathrm{Ma} \ll [1](/page/1). This neglects effects, such as , allowing the to simplify to \nabla \cdot \mathbf{u} = 0, where \mathbf{u} is the velocity field. The approximation assumes that inertial forces due to fluid acceleration are negligible compared to viscous forces, which holds when the \mathrm{Re} \ll [1](/page/1). A canonical example is , describing quasistatic creeping motion around objects like in viscous fluids at low Reynolds numbers. In this scenario, the Navier-Stokes equations reduce to the linear Stokes equations, \nabla [p](/page/Pressure) = \eta \nabla^2 \mathbf{u} and \nabla \cdot \mathbf{u} = 0, where p is and \eta is dynamic viscosity. For a of radius R moving with velocity U, the drag force is given by : F = 6 \pi \eta R U. This result, derived for steady translation, extends quasistatically to slowly varying motions where transient effects are minimal. Biological flows provide practical applications of this approximation. In blood circulation through capillaries, where vessel diameters are on the order of 5–10 \mum and flow speeds are below 1 mm/s, the is typically \mathrm{Re} < 0.1, enabling quasistatic modeling of red blood cell motion and plasma flow as Stokes-like creeping flows. Similarly, in insect flight, such as that of fruit flies or bees, the quasi-steady approximation captures wing kinematics during hovering or slow maneuvers, where flapping frequencies yield Strouhal numbers \mathrm{St} \approx 0.2–0.4. Here, aerodynamic forces are estimated by integrating translational and rotational contributions over the wing stroke, aligning well with experimental force measurements despite intermediate s (Re ≈ 100–10^4). Despite its utility, the quasistatic approximation has limitations in low-speed flows involving significant unsteadiness. It fails to capture phenomena like unsteady wakes or vortex shedding, which emerge when oscillatory motions introduce substantial temporal variations. The error increases with the Strouhal number \mathrm{St} = f L / U > 0.1, where f is oscillation frequency, L is a , and U is flow speed, as this parameter quantifies the ratio of local acceleration to convective acceleration; beyond this threshold, full unsteady Navier-Stokes solutions are required for accuracy.

Electromagnetic applications

Quasi-static fields

In electromagnetism, the quasistatic approximation applies to time-varying fields that evolve slowly enough to be treated as instantaneous static configurations, neglecting wave propagation effects. This regime is valid when the characteristic frequency ω of field variations satisfies ω ≪ c/L, where c is the speed of light and L is the typical system size, ensuring that the time for light to traverse the system is much shorter than the variation timescale. In the quasistatic approximation, depending on whether electric or magnetic effects dominate, either the displacement current term ∂D/∂t is omitted from Ampère's law (magnetoquasistatic case) or the magnetic induction term ∂B/∂t is neglected in Faraday's law (electroquasistatic case). This reduces Maxwell's equations to coupled time-dependent electrostatic and magnetostatic forms. In the electroquasistatic case, applicable to slowly varying charge distributions, the electric field satisfies ∇ · = ρ () and ∇ × = 0 (irrotational field), with = ε in linear media. These equations imply that the φ, defined by = -∇φ, obeys ∇ · (ε ∇φ) = -ρ for regions with ρ, or ∇²φ = 0 in charge-free spaces. This formulation allows the electric field to be computed as if charges were static at each instant, capturing capacitive effects in circuits without radiative losses. For the magnetoquasistatic case, relevant to slowly varying currents, the B obeys ∇ · B = 0 (no magnetic monopoles) and ∇ × H = J (Ampère's law without ), where H = B/μ in linear media and J is the . The A, satisfying B = ∇ × A in the Coulomb gauge ∇ · A = 0, enables solution via the equation ∇ × (μ⁻¹ ∇ × A) = J, often simplified to a Poisson-like form in uniform media. This approach accounts for inductive effects while ignoring propagation delays. A practical example occurs in low-frequency electrical circuits, such as those operating below the range, where the skin depth δ = √(2/ωμσ)—with μ the permeability and σ the —exceeds the dimensions. This condition ensures uniform distribution and field penetration, justifying uniform quasistatic fields inside components like inductors and allowing lumped circuit models to predict behavior accurately without full wave analysis.

Retarded time perspective

In the context of , the quasistatic approximation can be understood through the lens of retarded potentials, which represent the exact general solution to in the Lorenz gauge for prescribed charge and current densities. The at position \mathbf{r} and time t is expressed as \phi(\mathbf{r}, t) = \frac{1}{4\pi \epsilon_0} \int \frac{\rho(\mathbf{r}', t_r)}{|\mathbf{r} - \mathbf{r}'|} \, dV', where t_r = t - |\mathbf{r} - \mathbf{r}'|/c is the accounting for the finite propagation speed c of electromagnetic signals, and \rho(\mathbf{r}', t_r) is the evaluated at that delayed time. The follows analogously: \mathbf{A}(\mathbf{r}, t) = \frac{\mu_0}{4\pi} \int \frac{\mathbf{J}(\mathbf{r}', t_r)}{|\mathbf{r} - \mathbf{r}'|} \, dV', with \mathbf{J} the current density at the retarded time. These integrals incorporate the causal structure of electromagnetic interactions, where effects from sources at \mathbf{r}' influence the field at \mathbf{r} only after the light-travel delay |\mathbf{r} - \mathbf{r}'|/c. The quasistatic limit emerges when this retardation becomes negligible, specifically when the maximum light-travel time across the system, on the order of L/c for characteristic size L, is much smaller than the intrinsic time scale \tau over which the sources vary significantly (L/c \ll \tau). In this regime, t_r \approx t, allowing the potentials to simplify to instantaneous forms: \phi(\mathbf{r}, t) \approx \frac{1}{4\pi \epsilon_0} \int \frac{\rho(\mathbf{r}', t)}{|\mathbf{r} - \mathbf{r}'|} \, dV', and similarly for \mathbf{A}, effectively decoupling the fields from propagation delays and reducing the description to time-dependent but non-retarded integrals akin to static Coulomb and Biot-Savart laws. This viewpoint underscores the approximation's physical basis: it discards the finite-speed corrections inherent in full dynamic solutions, valid for scenarios where electromagnetic signals traverse the system nearly instantaneously relative to source evolution. For instance, in typical laboratory setups with dimensions L \approx 1 m operating at audio frequencies (f \approx 1 kHz, so \tau \approx 1 ms), the condition holds since c \tau \approx 3 \times 10^5 m \gg L. This perspective bridges the quasistatic approximation to more complete dynamic treatments, revealing when and wave propagation effects—arising from higher-order expansions of the —can be safely ignored. In antenna theory, it manifests as the near-field approximation, where for distances r \ll \lambda/2\pi (\lambda the ), the fields of a small radiating element like an electric dipole are dominated by quasistatic terms decaying as $1/r^3, neglecting the $1/r components that require full . This regime is crucial for applications such as or , where the system's compactness ensures minimal propagation delays.

Mathematical foundations

Scaling analysis

The quasistatic approximation relies on nondimensional to identify regimes where time-dependent effects can be neglected relative to dominant spatial or equilibrium processes, unifying criteria across , , and through small parameters that quantify timescale separations. In general, these parameters, denoted as ε, represent the ratio of a characteristic relaxation or timescale to the timescale of the external process, with the holding when ε ≪ 1. This approach transforms governing equations into dimensionless form, revealing the of terms involving time derivatives compared to steady-state contributions. In , the small parameter is typically ε = τ_relax / τ_process, where τ_relax is the system's internal relaxation time to (e.g., or equilibration) and τ_process is the duration of the imposed change. When ε ≪ 1, the system remains in near- throughout, allowing static thermodynamic relations to approximate the . Order-of-magnitude estimates show that time derivatives of variables scale as O(1/τ_process), while relaxation terms balance at O(1/τ_relax); thus, the approximation neglects transients if the evolves much slower than relaxation. This criterion ensures minimal beyond quasistatic limits in finite-time . In , of the Navier-Stokes equations introduces the St = L / (U τ_process) = f L / U, where f = 1/τ_process is the driving frequency, L is a , and U is a convective . The unsteady acceleration term scales as O(U / τ_process) = O(U² / L) · St, which is negligible compared to the convective term O(U² / L) when St ≪ 1, justifying the quasi-steady flow approximation. This holds for low-frequency oscillations where local acceleration is dwarfed by advection, as seen in airfoil pitching or scenarios at low reduced frequencies. In , the key parameter is ε = v / c or equivalently ε = ω L / c, with v the characteristic velocity of sources, ω the , L the system size, and c the . Time derivatives in scale as O(ω) for fields, while spatial gradients scale as O(1/L); the or retardation effects become small when ε ≪ 1, reducing to electroquasistatic or magnetoquasistatic limits. Dimensional analysis confirms that Faraday and Ampère terms balance without radiative corrections if the propagation time L/c ≪ τ_process. This scaling framework unifies the fields by highlighting a common structure: the quasistatic regime emerges when the dimensionless frequency parameter ε = ω L / U_char ≪ 1, where U_char is the characteristic speed (e.g., sound speed in fluids for acoustic limits, akin to c in electromagnetism, or equilibrium velocity in thermodynamics). In acoustic fluid flows, for instance, low St combined with low Mach number mirrors the EM condition, ensuring inertial or wave propagation terms do not dominate over quasistatic balances.

Derivation of approximation criteria

The quasistatic approximation is formalized through asymptotic expansions in perturbation theory, where a small parameter ε quantifies the slowness of temporal variations relative to spatial or other scales, allowing the neglect of time derivatives to leading order. Consider a general governing equation of the form ∂f/∂t = L, where L is a spatial operator and f represents the state variable (e.g., fields, densities, or thermodynamic potentials). Assume an asymptotic solution f = f₀ + ε f₁ + ε² f₂ + ..., with ε ≪ 1 representing the ratio of characteristic time scales, such as the rate of external driving to the system's intrinsic relaxation time. Substituting this expansion yields, at O(1): 0 = L[f₀], recovering the static equilibrium solution f₀. At O(ε): ∂f₀/∂t = L[f₁], where the time derivative of the leading-order solution drives the first correction, but higher-order terms remain small if ε is sufficiently small. This structure ensures that the quasistatic solution approximates the full dynamic behavior with controlled error, as validated in perturbation analyses of slow-varying systems. In , the quasistatic criterion emerges from the second law, dS ≥ đQ / T, with equality for . For a , the system remains in near-equilibrium, allowing the identification đQ ≈ T dS to hold exactly at leading order in the perturbation expansion. To derive this, expand the around a reference : dS = (đQ / T) + ε δ, where δ ≥ 0 captures dissipative contributions from finite-rate effects, and ε parameterizes the deviation from infinitely slow driving. The du = đQ - p dV implies that, to O(1), the process satisfies the reversible relation du = T dS - p dV, with zero (δ = 0). Higher-order terms introduce small irreversibilities, such as viscous heating or frictional losses, but these are O(ε) and negligible for slow processes, justifying the approximation đQ = T dS for computing changes. This first-order reversibility underpins applications like slowly compressed gases, where the error in calculation scales with the driving rate. In , the perturbation expansion applies to the unsteady Navier-Stokes equations, ∇ · u = 0 and ∂u/∂t + (u · ∇) u = -∇p/ρ + ν ∇² u, where the time derivative term is scaled by ε = τ / T, with τ the flow relaxation time and T the external variation time (T ≫ τ). Expanding u = u₀ + ε u₁ + ..., p = p₀ + ε p₁ + ..., the leading-order equations (O(1)) reduce to the steady incompressible Navier-Stokes: ∇ · u₀ = 0, (u₀ · ∇) u₀ = -∇p₀/ρ + ν ∇² u₀, treating the flow as quasi-steady despite slow parameter changes (e.g., varying or ). The O(ε) correction incorporates ∂u₀/∂t on the right-hand side, providing unsteady adjustments like or history effects in low-Reynolds-number flows. This expansion is valid when ε ≪ 1, as in creeping flows or at low reduced frequencies, where the steady-state solution dominates. Error bounds in the quasistatic approximation arise from the in the asymptotic series, typically O(ε) for the first correction and higher for subsequent terms. For instance, the neglected unsteady contributions, such as ∂u/∂t in fluids or finite-rate dissipation in , introduce relative errors bounded by ε times the magnitude of the leading-order , e.g., |f - f₀| ≤ C ε ||f₀|| for some constant C depending on the norms. These bounds confirm the approximation's accuracy for sufficiently slow processes, with higher-order expansions (e.g., including ε f₁) reducing the error to O(ε²).

References

  1. [1]
    3.3 Conditions for Fields to be Quasistatic - MIT
    Thus, either of the quasistatic approximations is valid if an electromagnetic wave can propagate a characteristic length of the system in a time that is short ...Missing: physics | Show results with:physics
  2. [2]
  3. [3]
    Electromagnetics from a quasistatic perspective - AIP Publishing
    Mar 1, 2007 · Quasistatic models provide intermediate levels of electromagnetic theory in between statics and the full set of Maxwell's equations.
  4. [4]
    Quasi-Static Processes - Richard Fitzpatrick
    A quasi-static process must be carried out on a timescale that is much longer than the relaxation time of the system.Missing: approximation | Show results with:approximation
  5. [5]
    3.4 Thermodynamic Processes – General Physics Using Calculus I
    A quasi-static process refers to an idealized or imagined process where the change in state is made infinitesimally slowly so that at each instant, the system ...
  6. [6]
    [PDF] Chapter 14 Quasi-static approximations ∇ · E = ρ /ε ∇ × E = − ∂B /∂t
    A quasi-static system means a system for which (i).. the sources and fields change slowly; (ii) .. and one or more of the red terms in the field equations can ...Missing: physics | Show results with:physics
  7. [7]
    [PDF] The Impossible Process: Thermodynamic Reversibility - CORE
    Aug 8, 2016 · This is the notion of thermodynamically reversible or quasi-static processes. They are used in the definition of entropy and are distinguished ...
  8. [8]
    The impossible process: Thermodynamic reversibility - ScienceDirect
    Thermodynamically reversible processes carry contradictory properties. They are supposed to be in equilibrium yet still change their state.
  9. [9]
  10. [10]
    [PDF] On the validity of the stochastic quasi-steady-state approximation in ...
    Abstract. The quasi-steady-state approximation is widely used to develop simplified deterministic or stochastic models of enzyme catalyzed reactions.Missing: relaxation | Show results with:relaxation
  11. [11]
    [PDF] 6 Quasistatic thermodynamic processes
    6.1 Definition of quasistatic processes. Quasistatic processes are processes that proceed slowly enough that the system is in internal equilibrium.Missing: approximation | Show results with:approximation
  12. [12]
    [PDF] Lecture 2 The First Law of Thermodynamics (Ch.1) - SMU Physics
    Summary of quasi-static processes of ideal gas. Quasi-Static process. ΔU. Q. W. Ideal gas law isobaric. (ΔP=0) isochoric. (ΔV=0). 0 isothermal. (ΔT=0). 0.
  13. [13]
    [PDF] 8.044 Lecture Notes Chapter 3: Thermodynamics, first pass
    Quasistatic processes: (slowness) All changes must be made so slowly that every state through which the system passes may be considered an equilibrium state. In ...
  14. [14]
  15. [15]
  16. [16]
    [PDF] Physics 115 Thermodynamic processes
    Apr 21, 2014 · – Reversible process for ideal gas, ideal reservoir, no friction ... • If ideal gas expands but no heat flows, work is done by gas (W>0 ...Missing: PdV | Show results with:PdV
  17. [17]
    Irreversible Adiabatic Compression of an Ideal Gas
    Suppose that instead of suddenly placing the weight on the piston, it is lowered on infinitesimally slowly (using a crane say). A quasistatic adiabatic.
  18. [18]
    4.3 Features of reversible processes - MIT
    A process must be quasi-static (quasi-equilibrium) to be reversible. This means that the following effects must be absent or negligible: Friction: If.Missing: equivalence | Show results with:equivalence
  19. [19]
    reversible and irreversible processes, entropy and introduction ... - MIT
    For reversible processes (the most efficient processes possible), the net change in entropy in the universe (system + surroundings) is zero. Phenomena that ...Missing: production | Show results with:production
  20. [20]
    PHYS 200 - The Second Law of Thermodynamics and Carnot's Engine
    You never deviate too much from equilibrium, and it's called a quasi-static; that means not quite static, but as slow as you like, and it is called reversible.
  21. [21]
    [PDF] Lecture 5: Thermodynamics
    ball rolling on the ground with friction is quasistatic but not reversible. Generally, any reversible thermodynamic process must be quasistatic. The ...
  22. [22]
    [PDF] Fluid Dynamics II (D)
    Sep 18, 2020 · • If Sr 1, then the flow is quasi-steady, and one can often ignore ... ary layer equations as an approximation to the Navier-Stokes equation, even.
  23. [23]
    [PDF] Approximate Solutions of the Navier-Stokes Equation - Moodle@Units
    If the characteristic frequency f is very small such that. St<<1, the flow is called quasi-steady. The effect of gravity is important only in flows with free- ...
  24. [24]
    The effect of uncertain bottom friction on estimates of tidal current ...
    Jan 9, 2019 · Neglecting the inertial term in (2.2) (by assumption of the quasi-steady limit) requires that the head difference driving the flow is balanced ...
  25. [25]
    The unsteady motion of solid bodies in creeping flows
    Apr 26, 2006 · In treating unsteady particle motions in creeping flows, a quasi-steady approximation is often used, which assumes that the particle's motion ...
  26. [26]
    Inertial Lubrication Theory | Phys. Rev. Lett.
    May 7, 2010 · In this highly dissipative condition, the dynamics of the fluid film is obtained from the quasistatic balance between viscous forces and ...
  27. [27]
    The mathematical theory of low Mach number flows
    Incompressible fluid flow differs from compressible fluid flow in that one of the equations of evolution is replaced by the constraint that the flow be ...
  28. [28]
    [PDF] Low Reynolds number hydrodynamics
    number, the better will be the approximate solution of the Navier-Stokes equations obtained by retaining only viscous terms. The exact Reynolds number above ...
  29. [29]
    Mathematical Models for Some Aspects of Blood Microcirculation
    In the present paper, we review some recent results in modeling blood flow in such small vessels, considering three areas.
  30. [30]
    The aerodynamics of hovering insect flight. I. The quasi-steady ...
    The conventional aerodynamic analysis of flapping animal flight invokes the 'quasisteady assumption' to reduce a problem in dynamics to a succession of static ...
  31. [31]
    Experimental study on the quasi-steady approximation of glottal flows
    May 10, 2022 · ... limitation in the oscillation range (227 and 195 Pa for lower ... The Strouhal number appears thus as an indicator of the importance ...
  32. [32]
    3 Introduction To Electroquasistatics and Magnetoquasistatics - MIT
    3.2 Quasistatic Laws. The quasistatic laws are obtained from Maxwell's equations by neglecting either the magnetic induction or the electric displacement ...
  33. [33]
    [PDF] Classical Electromagnetism - Richard Fitzpatrick
    The above potentials are termed retarded potentials (because the integrands are evaluated at the retarded time). Finally, according to the discussion in the ...
  34. [34]
    [PDF] On quasi-static models hidden in Maxwell's equations
    In this paper we introduce the electromagnetic quasi-static models in a simple but mean- ingful way, relying on the dimensional analysis of Maxwell's ...
  35. [35]
    [PDF] A Variational Principle for Quasistatic Mechanics - Michael Peshkin
    The quasistatic approximation is appropriate for many interesting driven dissipative systems below a characteristic driving velocity. Bounds on the error caused ...