Fact-checked by Grok 2 weeks ago

Hydrostatic equilibrium

Hydrostatic equilibrium is a fundamental concept in describing the condition in which the gravitational force acting on a element is precisely balanced by the , resulting in no and thus a static with zero . This balance is mathematically expressed by the , \frac{dP}{dz} = -\rho g, where P is , \rho is density, g is , and z is height, indicating that pressure decreases with altitude to counteract the weight of the overlying . In this state, the remains at rest relative to its container or , assuming no other forces like or dominate. The principle of hydrostatic equilibrium underpins numerous natural phenomena across scales in physics and . In planetary and stellar atmospheres, it governs the vertical distribution of and , enabling models of atmospheric where the weight of air above a given level is supported by the difference below. For instance, Earth's atmosphere maintains approximate hydrostatic balance, with decreasing exponentially with height due to the near-constant . In stellar interiors, hydrostatic equilibrium is essential for stable and evolution, where the inward pull of on the stellar material is opposed by outward from and radiation, preventing collapse or expansion over much of a star's lifetime. This equilibrium also applies to self-gravitating fluids in astrophysical contexts, such as the cores of giant planets or neutron , where deviations can lead to dynamical instabilities. Beyond gaseous systems, hydrostatic equilibrium influences liquid bodies like oceans and lakes on Earth, where it determines pressure profiles with depth and supports applications in engineering, such as dam design and submarine operations. In more extreme environments, it features in planetary geology and dwarf planet classification, distinguishing rounded bodies shaped by self-gravity from irregular ones. Overall, the concept provides a cornerstone for deriving equations of state and simulating fluid behaviors in both terrestrial and cosmic settings, with ongoing research refining its role in non-ideal fluids under high pressures or magnetic fields.

Fundamentals

Definition and Principles

Hydrostatic equilibrium describes the condition in a fluid or plastic solid where gravitational forces are precisely balanced by the pressure gradient, resulting in no net acceleration of fluid elements and maintaining a static configuration. This balance occurs when the downward pull of gravity on a fluid parcel is exactly counteracted by the upward force from the surrounding higher pressure below it, ensuring stability without motion. A fundamental principle is that pressure within the fluid increases with depth due to the accumulating weight of the overlying material, creating a vertical that supports the structure. This leads to the hydrostatic paradox, where the pressure at a given depth—and thus the force on a horizontal surface—is independent of the container's shape or the total volume of fluid above, depending solely on the height of the fluid column and its . The counterintuitive nature arises because the base force can exceed the fluid's total weight in non-cylindrical containers, but equilibrium is restored when considering the net forces on the container walls, which transmit the excess to the supports. Simple examples illustrate this concept: in a glass of , the pressure at the bottom supports the weight of the above, regardless of the glass's taper. Similarly, in Earth's atmosphere, air decreases with altitude as the weight of the air overhead diminishes, maintaining approximate hydrostatic balance under calm conditions. Key assumptions underpin these basic cases, including the fluid's incompressibility, where remains constant, and static conditions with no flow or viscous effects influencing the balance. These simplifications hold well for liquids like but approximate gaseous atmospheres, where varies.

Historical Context

The concept of hydrostatic equilibrium traces its roots to ancient observations of and , most notably articulated by around 250 BCE in his work . states that the buoyant force on an object immersed in a equals the weight of the displaced, laying the groundwork for understanding variations in fluids under gravity. This insight emerged from experiments, such as determining the purity of a gold crown by measuring displacement, and represented an early recognition that fluids exert upward forces proportional to submerged volume, essential for later equilibrium concepts. In the , advancements in experimental built on these foundations, with inventing the mercury in 1643, which demonstrated that supports a column of mercury about 760 mm high and refuted the notion of a perfect vacuum's "horror." This device quantified air pressure's role in balance, showing how external forces maintain equilibrium in confined liquids. Concurrently, explored pressure transmission in the 1650s through experiments like the demonstration, culminating in his 1663 publication Traité de l'équilibre des liqueurs, which formalized that pressure applied to a confined propagates undiminished in all directions, enabling precise analysis of hydrostatic balance in vessels. The 18th and 19th centuries saw hydrostatic equilibrium integrated into gravitational theories, with introducing the for gravitational fields in 1777 to describe force distributions in continuous media like fluids. advanced this in the 1780s by deriving the Poisson equation for gravitational potentials, applying it to fluid stability in and demonstrating how density variations maintain under self-gravity. further bridged with microscopic dynamics in the 1870s, deriving the via kinetic theory to explain the exponential density decrease with height in isothermal atmospheres, thus reconciling macroscopic pressure gradients with molecular motion. Twentieth-century extensions expanded hydrostatic equilibrium to relativistic and astrophysical realms. 's equivalence principle, formulated in 1907 and central to his 1915 general , equated gravitational fields with accelerated frames, generalizing hydrostatic balance to curved where pressure gradients counter . In the 1920s, applied these ideas to stellar interiors in works like The Internal Constitution of the Stars (1926), using hydrostatic equilibrium alongside to model energy transport and stability in stars, predicting mass-luminosity relations that explained observed stellar diversity.

Mathematical Formulation

Derivation from Force Balance

Hydrostatic equilibrium arises from the balance of forces acting on a small of within a under non-relativistic conditions. Consider a thin, cylindrical with cross-sectional area A and height \Delta z, oriented vertically in a uniform where the acceleration due to gravity is g directed downward. The forces on this parcel include the pressure forces acting on its top and bottom surfaces and the gravitational force due to its own weight. The pressure at the bottom surface (at height z) exerts an upward force of P(z) A, while the pressure at the top surface (at height z + \Delta z) exerts a downward force of P(z + \Delta z) A. The net force from the pressure difference is [P(z) - P(z + \Delta z)] A, which points upward if pressure decreases with height. The downward gravitational force, or weight of the parcel, is \rho g A \Delta z, where \rho is the fluid density. In hydrostatic equilibrium, the parcel experiences no net acceleration, so the forces balance: [P(z) - P(z + \Delta z)] A = \rho g A \Delta z. Dividing by A \Delta z and taking the limit as \Delta z \to 0 yields the differential form \frac{dP}{dz} = -\rho g, where the negative sign indicates pressure decreases upward. For a fluid of constant density, integrating \frac{dP}{dz} = -\rho g from height z = 0 (where P = P_0) to arbitrary z gives the barometric formula P(z) = P_0 - \rho g z. This linear pressure profile holds under the assumptions of one-dimensional variation along the vertical (z) direction, static conditions with no fluid motion, and constant gravitational acceleration g. These assumptions limit the applicability of the simple derivation; for instance, it does not account for density variations in compressible fluids, where \rho depends on P, requiring numerical or more advanced solutions. Similarly, it fails in relativistic contexts involving strong gravitational fields. In a more general setting with non-uniform , the equation extends to vector form by considering the balance between the and the gravitational force per unit volume: \nabla P = -\rho \nabla \Phi, where \Phi is the satisfying \mathbf{g} = -\nabla \Phi. This form applies to arbitrary gravitational fields while maintaining the hydrostatic (no motion) condition.

Derivation from Navier-Stokes Equations

The Navier-Stokes equations describe the motion of viscous fluids, providing a fundamental framework for understanding fluid dynamics. The momentum equation in its general form for a Newtonian fluid is given by \rho \left( \frac{\partial \mathbf{v}}{\partial t} + \mathbf{v} \cdot \nabla \mathbf{v} \right) = -\nabla p + \mu \nabla^2 \mathbf{v} + \left( \mu + \lambda \right) \nabla (\nabla \cdot \mathbf{v}) + \rho \mathbf{g}, where \rho is the fluid density, \mathbf{v} is the velocity field, p is the pressure, \mu is the dynamic viscosity, \lambda is the second viscosity coefficient (often related to bulk viscosity), and \mathbf{g} is the gravitational acceleration vector. This equation arises from Newton's second law applied to a fluid element, balancing inertial forces with pressure gradients, viscous stresses, and body forces like gravity. To derive hydrostatic equilibrium, consider the static limit where the fluid is at rest, so \mathbf{v} = \mathbf{0} everywhere and there are no time-dependent changes (\partial / \partial t = 0). Substituting these conditions into the Navier-Stokes momentum equation yields $0 = -\nabla p + \mu \nabla^2 (\mathbf{0}) + \left( \mu + \lambda \right) \nabla (\nabla \cdot \mathbf{0}) + \rho \mathbf{g}, which simplifies immediately to \nabla p = \rho \mathbf{g}. This is the vector form of the hydrostatic equilibrium equation, indicating that the pressure gradient precisely balances the gravitational body force per unit volume. The viscous terms vanish because they depend on velocity gradients, which are zero in a static fluid; specifically, the shear stress tensor components, proportional to \partial v_i / \partial x_j + \partial v_j / \partial x_i, become zero, leaving no contribution from viscosity. For fluids in hydrostatic equilibrium, the derivation holds whether the fluid is incompressible (\rho = constant) or compressible (\rho varies with position or ). In the incompressible case, integrating \nabla p = \rho \mathbf{g} along the direction of (assuming \mathbf{g} = -g \hat{z}) gives a linear pressure profile p(z) = p_0 - \rho g z. For compressible fluids, such as gases, \rho may depend on p via an (e.g., ), leading to more complex profiles solved iteratively or numerically. In the broader context of , hydrostatic equilibrium represents a steady-state to the Navier-Stokes equations, but small around this can excite dynamic modes. For instance, introducing a \delta \mathbf{v} leads to restorative forces from the , resulting in oscillations such as or buoyancy-driven instabilities, which highlight how the equilibrium acts as a for more general flows.

Derivation in General Relativity

In , the derivation of hydrostatic equilibrium for a static, spherically symmetric configuration of begins with the in : ds^2 = -e^{2\Phi(r)} \, dt^2 + e^{2\Lambda(r)} \, dr^2 + r^2 \, d\Omega^2, where \Phi(r) and \Lambda(r) are functions depending on the radial coordinate r, and d\Omega^2 = d\theta^2 + \sin^2\theta \, d\phi^2. This assumes time-independence and spherical symmetry, suitable for equilibrium states without or other . The matter is modeled as a perfect fluid with stress-energy tensor T^{\mu\nu} = (\rho + P) u^\mu u^\nu + P \, g^{\mu\nu}, where \rho is the total energy density (including rest mass), P is the isotropic pressure, u^\mu is the four-velocity normalized such that u^\mu u_\mu = -1, and g^{\mu\nu} is the inverse metric tensor. For a static fluid at rest in these coordinates, u^\mu = (e^{-\Phi}, 0, 0, 0), so the non-zero components are T^{tt} = \rho e^{-2\Phi}, T^{rr} = P e^{-2\Lambda}, T^{\theta\theta} = P / r^2, and T^{\phi\phi} = P / (r^2 \sin^2 \theta). The relate geometry to matter via G_{\mu\nu} = 8\pi T_{\mu\nu} (in units where G = c = 1), where G_{\mu\nu} is the . Hydrostatic equilibrium is enforced by the local and momentum, \nabla_\mu T^{\mu\nu} = 0. In the comoving frame, this condition implies no , and for the radial component (\nu = r), it yields a balance between the and gravitational effects encoded in the . Solving the field equations for the tt and rr components gives expressions for the functions: e^{-2\Lambda(r)} = 1 - 2m(r)/r, where m(r) = \int_0^r 4\pi s^2 \rho(s) \, ds is the gravitational mass enclosed within radius r, and \frac{d\Phi}{dr} = \frac{m(r) + 4\pi r^3 P}{r^2 (1 - 2m(r)/r)}. Substituting into the radial conservation equation \nabla_\mu T^{\mu r} = 0 produces the Tolman-Oppenheimer-Volkoff (TOV) equation: \frac{dP}{dr} = -(\rho + P) \frac{m(r) + 4\pi r^3 P}{r^2 \left(1 - \frac{2m(r)}{r}\right)}. This differential equation, first obtained by Tolman and independently by Oppenheimer and Volkoff, couples pressure and density gradients to the enclosed mass and relativistic corrections from spacetime curvature. In the weak-field, non-relativistic limit (where \rho \ll 1, P \ll 1, and m(r)/r \ll 1 in these units, equivalent to small G), the TOV equation simplifies by neglecting terms like P compared to \rho and the $4\pi r^3 P correction, recovering the Newtonian hydrostatic balance \frac{dP}{dr} = -\rho \frac{m(r)}{r^2}. The TOV equation requires an P = P(\rho) to close the system and is typically solved numerically via outward from the stellar , where m(0) = 0, P(0) and \rho(0) are specified central values, and \frac{dP}{dr}\big|_{r=0} = 0 by , continuing until P(r) = 0 defines the surface radius. This framework is vital for describing compact objects such as white dwarfs, neutron stars, and quark stars, where strong demands relativistic treatment.

Applications

Fluid Statics

Fluid statics concerns the behavior of at rest under the influence of , where hydrostatic equilibrium ensures that the balances the weight of the fluid, preventing any net motion. In incompressible like , this balance leads to a linear increase in with depth, fundamental to numerous designs. The P at a depth h in a of constant \rho under g is given by P = \rho g h, assuming the pressure at the surface is negligible or accounted for separately. This formula arises from the balance in hydrostatic equilibrium and is applied in calculating forces on submerged structures. For instance, in dam design, the hydrostatic pressure on the upstream face determines the required thickness to resist overturning and sliding. must withstand hull stresses from this pressure, which increases by about 1 atm every 10 meters of depth, necessitating robust materials like high-strength . Hydraulic presses exploit Pascal's principle, derived from hydrostatic equilibrium, where a small input on a confined produces a proportionally larger output on a larger , enabling heavy lifting in industrial applications. For compressible fluids, such as gases in the atmosphere, density varies with pressure, leading to an exponential decay. In an isothermal ideal gas, the barometric formula describes the pressure P(z) at height z as P(z) = P_0 \exp\left(-\frac{M g z}{R T}\right), where P_0 is the surface pressure, M is the molar mass, R is the gas constant, and T is the temperature. This arises from integrating the hydrostatic equilibrium equation with the ideal gas law and applies to Earth's lower atmosphere, explaining why air pressure halves roughly every 5.5 km. Devices like manometers and barometers rely on hydrostatic equilibrium to measure . A manometer, filled with a such as mercury, equates the pressure difference to the difference h of the liquid columns via \Delta P = \rho g h, allowing precise readings in pipes or vessels. Barometers, often using a mercury column in a closed tube, measure absolute by balancing it against the fluid's weight, with standard sea-level pressure supporting a 760 mm column. In setups with different fluids, such as over , the levels adjust until pressures equalize, demonstrating immiscible fluid . Hydrostatic equilibrium can be disrupted by instabilities, notably the Rayleigh-Taylor instability, which occurs when a denser fluid overlies a lighter one under , causing the interface to deform into spikes and bubbles that promote mixing. This qualitative behavior arises from perturbations growing due to unbalanced gravitational forces at the interface. In , hydrostatic equilibrium informs safety factors for fluid containment to prevent failures from buildup. Dams and reservoirs incorporate uplift resistance with minimum safety factors of 1.4 against hydrostatic forces to account for uncertainties in loading. At extreme depths, such as the Mariana Trench's (about 11 km), hydrostatic reaches approximately 1000 (110 ), over 1000 times , highlighting the need for specialized submersibles with thick pressure hulls.

Astrophysics

In astrophysics, hydrostatic equilibrium is fundamental to understanding the internal structure of self-gravitating celestial bodies, particularly , where the inward pull of is balanced by outward gradients to maintain . This balance governs the distribution of , , and throughout a star's interior, enabling sustained in the core. For stars modeled as spherically symmetric, the principle integrates with equations of mass continuity and energy transport to predict observable properties like and . A key mathematical tool for modeling stellar interiors under hydrostatic equilibrium is the polytropic approximation, which assumes a relation between pressure and density of the form P = K \rho^{1 + 1/n}, where K is a constant and n is the polytropic index. This leads to the Lane-Emden equation, a second-order differential equation describing the structure of such polytropes: \frac{d}{dr} \left( r^2 \frac{d\theta}{dr} \right) = -r^2 \theta^n, with boundary conditions \theta(0) = 1 and d\theta/dr|_{r=0} = 0, where \theta is a dimensionless temperature or density variable, and r is scaled radially. Solutions to this equation, obtained numerically for most n, provide density and pressure profiles that approximate real stars, linking the equation of state to gravitational stability; for example, n = 3 models radiative interiors like those in massive stars. In the model, hydrostatic equilibrium determines the radial profiles of and , with reaching a central of approximately $2.5 \times 10^{16} to counteract gravitational . This is crucial for stability, as the high central (\sim 150 g/cm³) and temperature (\sim 15 million K) enable proton-proton chain reactions, while the ensures remains stable against collapse or expansion over billions of years. Standard solar models, solved iteratively with opacity and generation rates, confirm that deviations from would disrupt , but feedback mechanisms like increased opacity restore balance. For compact objects like white dwarfs and neutron stars, hydrostatic equilibrium incorporates degeneracy pressure to resist gravity. In white dwarfs, electron degeneracy pressure supports masses up to the of approximately 1.4 solar masses, beyond which relativistic effects cause instability, leading to collapse; this limit arises from integrating hydrostatic balance with the degenerate equation of state in the relativistic regime. Neutron stars, supported by neutron degeneracy pressure, achieve higher masses (up to ~2-3 solar masses) under the Tolman-Oppenheimer-Volkoff equation, a relativistic generalization of hydrostatic equilibrium that accounts for strong gravity. During stellar evolution, hydrostatic equilibrium is maintained throughout the main-sequence phase, where core hydrogen fusion provides the thermal pressure needed for balance, sustaining a star's structure for up to 10 billion years for solar-mass objects. Perturbations, such as radial pulsations, represent small oscillations around this equilibrium, analyzed via linearized ; these occur in evolved stars like Cepheids, where instability strips lead to periodic expansion and contraction without disrupting overall stability. Early models by Eddington in laid the groundwork for these concepts, using polytropes to explore radiative and convective equilibria in stellar interiors.

Planetary and Atmospheric Science

In planetary and atmospheric science, hydrostatic equilibrium governs the vertical structure of atmospheres, where the downward force of gravity is balanced by the upward pressure gradient. The fundamental equation describing this balance is \frac{dP}{dz} = -\rho g, where P is pressure, z is altitude, \rho is density, and g is gravitational acceleration. Integrating this with the ideal gas law yields the atmospheric scale height H = \frac{kT}{m g}, or equivalently H = \frac{RT}{M g} using molar quantities, which represents the altitude over which pressure decreases by a factor of e. For Earth's troposphere, with an average temperature of 250 K and molar mass M = 0.029 kg/mol, the scale height is approximately 7.3 km, leading to an exponential pressure drop where surface pressure (about 1013 hPa) falls to roughly 370 hPa at 7.3 km altitude. This profile is modulated by the temperature lapse rate, which affects density and thus the rate of pressure decrease. Representative examples illustrate these principles across solar system bodies. On , the tropospheric pressure drops by about two-thirds every 7 km due to this equilibrium, enabling patterns confined to the lower atmosphere. exemplifies an extreme case, with its thick atmosphere (96.5% CO₂) maintaining hydrostatic balance under surface pressures of approximately 92 —over 90 times 's—resulting in a of around 15-20 km despite high temperatures. This dense envelope traps heat via the while sustaining vertical stability. Saturn's moon provides another case, where its nitrogen-methane atmosphere, laden with organic particles, achieves hydrostatic equilibrium that supports hazy layers extending to 300 km altitude; this structure influences formation through meridional circulation and seasonal forcing, with polar clouds persisting for years post-solstice. Hydrostatic equilibrium also shapes planetary interiors, particularly in fluid-dominated regions like cores and mantles. Pressure gradients increase radially inward, with the core-mantle boundary (CMB) on experiencing about 136 GPa due to overlying mantle weight in hydrostatic balance. In planets with liquid outer cores, such as , these gradients drive convective motions in the molten iron-nickel alloy, which, combined with planetary rotation, generate dynamo effects producing magnetic fields that shield atmospheres from . This process requires sustained under hydrostatic constraints to maintain the geodynamo over billions of years. Rotation introduces centrifugal forces that modify effective gravity, given by g_{\text{eff}} = g - \Omega^2 r \cos^2 \phi, where \Omega is , r is radial , and \phi is ; this term peaks at the , reducing g_{\text{eff}} by up to 0.3% on . Consequently, rotating planets deviate from spherical shapes, forming oblate spheroids where equipotential surfaces align perpendicular to g_{\text{eff}}, flattening poles and bulging equators as seen in and . In hydrostatic equilibrium, this oblateness minimizes , with Earth's equatorial radius exceeding polar by 21 km. For exoplanets, hydrostatic models integrate these principles to assess , particularly for worlds with hydrogen-rich atmospheres like Hycean planets (hybrid water-worlds with ocean surfaces). These models solve hydrostatic equations alongside to predict pressure-temperature profiles, revealing habitable conditions (surface pressures 1-1000 bar, temperatures 273-395 K) on planets up to 10 masses within extended habitable zones. Such frameworks, applied to candidates like K2-18 b, evaluate detectability by balancing gravitational compression with atmospheric retention. Recent observations of K2-18 b, as of 2025, have reported tentative detections of potential biosignatures like , though these remain controversial and unconfirmed, highlighting the role of hydrostatic models in interpreting such data.

Geological Contexts

In geological contexts, hydrostatic equilibrium manifests as lithostatic pressure within Earth's solid interior, representing the vertical stress due to the weight of overlying rock layers. This pressure at a depth z is calculated as P(z) = \int_0^z \rho(z') g \, dz', where \rho(z') is the at depth z', and g is , accounting for the layered structure of the Earth with varying densities. The planet's average is approximately 5.51 g/cm³, reflecting the transition from lighter crustal rocks to denser and materials. In the mantle and core, equilibrium is maintained through viscoelastic behavior in solids, where rocks deform slowly over geological timescales via creep mechanisms, approximating hydrostatic conditions despite their rigidity on short scales. The mantle, behaving as a viscoelastic solid, supports minor deviatoric stresses that relax through diffusion creep or dislocation processes, allowing long-term isostatic adjustment. Isostasy exemplifies this balance, akin to , where the continental crust "floats" on the denser at greater depths than , explaining the elevation differences between continents and basins. Lithostatic pressure influences volcanism and tectonics by building up stress in magmatic systems, particularly in subduction zones where descending slabs increase overburden on the mantle wedge, promoting partial melting and volatile release. When magmatic pressure exceeds lithostatic levels—often by 10-50 MPa due to volatile exsolution—dykes propagate, leading to eruptions as the excess pressure overcomes rock tensile strength. In subduction settings, this pressure gradient drives fluid migration from the slab, lowering the solidus and facilitating arc volcanism. Density profiles supporting lithostatic equilibrium are inferred from seismic waves, whose velocities increase with depth due to rising and , revealing discontinuities like the core-mantle boundary at approximately 136 GPa. P- and S-wave data from global networks model these profiles, confirming mantle densities around 3.3-5.6 g/cm³ and core densities exceeding 9.9 g/cm³. Unlike fluids, where strictly prohibits shear stresses, rocks in Earth's interior achieve approximate hydrostatic balance through time-dependent and , enabling solids to support transient deviatoric loads before relaxing via mechanisms like pressure-solution creep in the upper crust. This viscoelastic response distinguishes geological , allowing tectonic deformation without immediate failure.

References

  1. [1]
    Fluid Statics & the Hydrostatic Equation – Introduction to Aerospace ...
    A fluid in static equilibrium is one in which every fluid particle is either at rest or has no relative motion with respect to the other particles in the fluid.
  2. [2]
    Hydrostatic equilibrium of the atmosphere - Richard Fitzpatrick
    Hydrostatic equilibrium occurs when the upward force from gas pressure balances the downward force from the weight of the gas slice. The equation of  ...
  3. [3]
    2.2 The Atmosphere's Pressure Structure: Hydrostatic Equilibrium
    Hydrostatic equilibrium is the balance of three forces: pressure from above, pressure from below, and the weight of the air parcel, resulting in no net force.
  4. [4]
    Stars - Imagine the Universe! - NASA
    This pressure counteracts the force of gravity, putting the star into what is called hydrostatic equilibrium. A star is okay as long as the star has this ...Missing: physics | Show results with:physics
  5. [5]
    [PDF] Hydrostatic Equilibrium - NMSU Astronomy
    Hydrostatic equilibrium is when gravitational force is balanced by outward pressure, where dFg + dFP = 0.Missing: physics | Show results with:physics
  6. [6]
    [PDF] The Hydrostatic Equation
    If the net upward pressure force on the slab is equal to the downward force of gravity on the slab, the atmosphere is said to be in hydrostatic balance. Page 4 ...
  7. [7]
    [PDF] Part B : Gas Physics
    If, furthermore, the gas starts off at rest (i.e. in a static configuration), the resulting equilibrium is called hydrostatic equilibrium.
  8. [8]
    Hydrostatic Equilibrium and the Homogeneous Atmosphere
    The principle of hydrostatic equilibrium is that the pressure at any point in a fluid at rest (whence, “hydrostatic”) is just due to the weight of the ...
  9. [9]
    [PDF] A simple explanation of the classic hydrostatic paradox
    The hydrostatic paradox states that in different shaped containers with the same base area, the liquid's force on the base is the same, but the liquid's weight ...
  10. [10]
    Archimedes Principle - an overview | ScienceDirect Topics
    The above statement is known as Archimedes' principle. One legend has it that Archimedes (Greek, lived in Syracuse—Sicily—between 287 and 212 BC) discovered ...
  11. [11]
    Torricelli and the Ocean of Air: The First Measurement of Barometric ...
    Torricelli was the first to make a mercury barometer and understand that the mercury was supported by the pressure of the air.
  12. [12]
    Traitez de l'équilibre des liqueurs et de la pesanteur de la masse de ...
    Dec 3, 2014 · Traitez de l'équilibre des liqueurs et de la pesanteur de la masse de l'air. by: Blaise Pascal. Publication date: 1663. Usage: Public Domain ...
  13. [13]
    Joseph-Louis Lagrange - Biography, Facts and Pictures
    May 30, 2018 · Created the concept of potential: the gravitational field, for example, is a potential field. ... Letter to Pierre Laplace, September 1782.
  14. [14]
    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 ...Missing: Louis | Show results with:Louis
  15. [15]
    (PDF) Arthur Stanley Eddington: pioneer of stellar structure theory
    Aug 7, 2025 · In 1920, Eddington pointed to the fusion of four hydrogen atoms into a helium atom as the likely energy source supplying the observed stellar luminosity.
  16. [16]
    [PDF] Hydrostatic Equilibrium and Exponential Atmospheres
    4. Going back to the original equation of hydrostatic equilibrium, dP/dr = −ρg, you see that the pressure increases more rapidly when the density is higher. ...
  17. [17]
    [PDF] lecture 8: the formation of bound structures, hydrostatic equilibrium ...
    Feb 27, 2008 · The lecture covers the formation of bound structures, hydrostatic equilibrium, and dark matter galaxy formation. It also discusses collapse and ...<|control11|><|separator|>
  18. [18]
    [PDF] The Momentum Conservation Equation Hydrostatic Equilibrium
    In words, the outward force due to the pressure gradient must balance the inward force due to gravity. Fig. 2 demonstrates hydrostatic equilibrium in action ...
  19. [19]
    [PDF] An introduction to fluid dynamics
    which is the equation of hydrostatic equilibrium. This equation tells us that ... This is called the Navier-Stokes Equation. Sometimes the. “continuity ...
  20. [20]
    [PDF] Interstellar Hydrodynamics 1. Introduction - Caltech (Tapir)
    Feb 23, 2011 · The above equations are called the Navier-Stokes ... Thus we see that the gas obeys the usual hydrostatic equilibrium equation, but with a ...<|separator|>
  21. [21]
  22. [22]
  23. [23]
    14.1 Fluids, Density, and Pressure – University Physics Volume 1
    Pressure due to the weight of a liquid of constant density is given by p = ρ g h , where p is the pressure, h is the depth of the liquid, ρ is the density of ...Missing: hydrostatic | Show results with:hydrostatic
  24. [24]
    [PDF] PART I FLUID DYNAMICS
    Figure 1-5. Hydrostatic pressure in a column of water. It is just a matter of multiplying the weight per unit volume of the liquid, ρg, by the volume of liquid ...Missing: formula source
  25. [25]
    [PDF] DESIGN OF GRAVITY DAMS - Bureau of Reclamation
    It serves as a guide to sound engineering practices in the design of concrete gravity dams and provides the technically trained, qualified design engineer with ...
  26. [26]
    [PDF] Physics, Chapter 8: Hydrostatics (Fluids at Rest)
    Whenever the pressure in a confined liquid is increased or diminished at any point, this change in pressure is transmitted equally throughout the entire liquid.
  27. [27]
    14.2 Measuring Pressure – University Physics Volume 1
    Manometers typically use a U-shaped tube of a fluid (often mercury) to measure pressure. A barometer (see Figure) is a device that typically uses a single ...
  28. [28]
    [PDF] A Physical Introduction to Fluid Mechanics - UW Courses Web Server
    Dec 19, 2013 · A hydraulic press uses the transmissibility of fluid pressure to produce large forces. ... This is a case where the force due to hydrostatic ...
  29. [29]
    [PDF] THEORY OF THE RAYLEIGH-TAYLOR INSTABILITY
    The instability of a heavy fluid layer supported by a light one is generally known as Rayleigh-Taylor (RT) instability. It can occur under gravity and, ...
  30. [30]
    [PDF] Guidelines for Design of Dams - NY.gov
    Hydrostatic uplift pressure from reservoir water and tailwater act on the dam. ... safety factors must be computed using the Friction Factor of Safety. The ...
  31. [31]
    Planet Postcard: The Mariana Trench | News
    Oct 17, 2017 · Part of the reason that the Challenger Deep is largely unexplored is due to hydrostatic pressure. When you enter any body of water and begin ...
  32. [32]
    [PDF] POLYTROPES
    This is known as the Lane-Emden equation for polytropic stars. The two boundary conditions for the eq. (poly.6) are at the center: θ = 1, dθ dξ. = 0, at ξ ...
  33. [33]
    [PDF] Polytropes – Derivation and Solutions of the Lane-Emden Equation
    Polytropes are useful as they provide simple solutions (albeit in some cases via numerical integration) for the internal structure of a star that can be ...
  34. [34]
    [PDF] Chapter 3 - Princeton University
    Thus, without pressure support, the Sun would collapse to a point within half an hour. This has not happened because the Sun is in hydrostatic equilibrium.
  35. [35]
    Fusion Regulation in the Sun - Stanford University
    Mar 7, 2011 · This relation gives us a direct relationship between pressure and density and allows us to solve for density and pressure profiles of the sun.
  36. [36]
    The Chandrasekhar limit: a simplified approach - IOPscience
    Jun 15, 2023 · This article outlines a simplified approach to approximating the Chandrasekhar limit for white dwarf stars at a level appropriate for advanced high school ...Abstract · Introduction · Relativistic degeneracy pressure · The Chandrasekhar mass
  37. [37]
    The Maximum Mass of Ideal White Dwarfs - ADS
    This mass (=0.910) is interpreted as representing the upper limit to the mass of an ideal white dwarf.
  38. [38]
    [PDF] STELLAR STRUCTURE AND EVOLUTION
    The relation (4.25) was derived by Arthur Eddington in the 1920s for his famous 'standard model'. He found that in regions with a high opacity κ (see Ch. 5) ...
  39. [39]
    [PDF] Radial Stellar Pulsations - Princeton University
    Up to this point, you have studied stars in hydrostatic equilibrium. In the next few lectures, we will consider oscillating stars. We begin with the ...
  40. [40]
    The Internal Constitution of the Stars - ADS - Astrophysics Data System
    Eddington, A. S.. Abstract. Foreword; Preface; 1. Survey of the problem; 2. Thermodynamics of radiation; 3. Quantum theory; 4. Polytropic gas spheres; 5 ...Missing: models | Show results with:models
  41. [41]
    None
    Nothing is retrieved...<|separator|>
  42. [42]
    The dynamics behind Titan's methane clouds - PNAS
    We present results of an axisymmetric global circulation model of Titan with a simplified suite of atmospheric physics forced by seasonally varying insolation.
  43. [43]
    Extensive iron–water exchange at Earth's core–mantle boundary ...
    Oct 15, 2024 · Chemical equilibrium between ferropericlase and molten iron to 134 GPa and implications for iron content at the bottom of the mantle.
  44. [44]
    Super‐Earth Internal Structures and Initial Thermal States - Boujibar
    Mar 31, 2020 · The generation and maintenance of a planetary magnetic field requires the presence of a large volume of electrically conductive fluid, large- ...
  45. [45]
  46. [46]
    How Newton Derived Shape of Earth | American Physical Society
    Oct 13, 2022 · He concluded that Earth, modeled as a homogeneous spheroid that rotates with uniform angular velocity, must have polar and equatorial axes with ...
  47. [47]
    Habitability and Biosignatures of Hycean Worlds - IOPscience
    Aug 26, 2021 · The model computes line-by-line radiative transfer in transmission geometry assuming a plane-parallel atmosphere in hydrostatic equilibrium.
  48. [48]
    Lithostatic Pressure - an overview | ScienceDirect Topics
    Lithostatic pressure is a function of rock density and is generally between 22.7 and 25.0 KPa/m (1.0 and 1.1 psi/ft) in coal-bearing successions (McKee et al., ...
  49. [49]
    Surface loading of a viscoelastic earth—I. General theory - Free
    In an earth model in hydrostatic equilibrium this unit vector is perpendicular to surfaces of constant pressure, density and gravitational potential.
  50. [50]
    Magmatic overpressures, volatile exsolution and potential ... - Nature
    Jun 10, 2020 · Magmatic overpressure is the main driving force of magma discharge through a conduit. The explosivity of an eruption is, in the simplest case, ...
  51. [51]
    Fluid flow in compressive tectonic settings: Implications for ...
    Apr 14, 2004 · The paradox posed by the conditions required to maintain high fluid pressure simultaneously with lower crustal rock strength can be explained by ...Missing: plasticity | Show results with:plasticity
  52. [52]
    Seismic Evidence for Internal Earth Structure
    Seismic velocities tend to gradually increase with depth in the mantle due to the increasing pressure, and therefore density, with depth. However, seismic ...
  53. [53]
    [PDF] The role of pressure solution creep in the ductility of the Earth's ...
    The aim of this review is to characterize the role of pressure solution creep in the ductility of the. Earth's upper crust and to describe how this creep ...
  54. [54]
    Does non-hydrostatic stress influence the equilibrium of ...
    For solid/solid systems the shift in equilibrium conditions for non-hydrostatic stresses can, in principle, be very large but such effects can be mitigated by ...