Fact-checked by Grok 2 weeks ago

Pressure-gradient force

The pressure-gradient force (PGF) is the net force per unit mass exerted on a due to spatial variations in , directing the parcel from regions of higher toward lower . This force arises from differences in molecular collisions across the parcel's surfaces, with its magnitude inversely proportional to and directly proportional to the steepness of the . In mathematical terms, the PGF is expressed as \mathbf{F}_{PGF} = -\frac{1}{\rho} \nabla P, where \rho is the and \nabla P is the vector; for the horizontal x-component, it simplifies to F_x / m = -\frac{1}{\rho} \frac{\partial P}{\partial x}. In atmospheric and oceanic dynamics, the PGF serves as the primary driver of large-scale fluid motions, such as and currents, by initiating along pressure gradients. Horizontally, it tends to produce straight-line flow perpendicular to isobars, with stronger gradients—indicated by closely spaced isobars on maps—resulting in higher speeds and more intense storms. Vertically, the PGF balances gravitational forces in , where pressure decreases with height at a rate of approximately \frac{\Delta P}{\Delta z} = -\rho g, with g being , preventing unrestricted upward motion in stable atmospheres. The PGF interacts with other forces, notably the Coriolis effect due to Earth's rotation, which deflects motion to the right in the Northern Hemisphere and left in the Southern Hemisphere, leading to geostrophic balance where winds flow parallel to isobars. In the absence of friction, this balance governs steady-state flows like trade winds and jet streams, while frictional effects near the surface cause cross-isobar flow toward low pressure. Pressure gradients themselves stem from variations in temperature, density, and altitude, making the PGF central to weather patterns and climate phenomena.

Fundamentals

Definition

The pressure-gradient force is the per unit mass exerted on a element arising from spatial variations in , directed from regions of to regions of low . This force acts perpendicular to surfaces of constant (isobars or isopycnals) and is proportional to the magnitude of the , representing the net effect of molecular collisions that imbalance across the . In , the pressure-gradient force serves as the primary mechanism that initiates and drives motion in the absence of viscous effects or other body forces, accelerating parcels toward lower pressure areas to equalize imbalances. Its magnitude corresponds to an in meters per second squared (m/s²), though the underlying has units of force per unit volume, such as pascals per meter (Pa/m) or newtons per cubic meter (N/m³); dividing by the modulates the resulting motion. thus influences the magnitude of the pressure-gradient force for a given . A conceptual illustration of this force occurs in a balloon releasing air, where the higher internal pressure creates a gradient across the opening, driving the air outward and propelling the balloon in the opposite direction via the reaction to this pressure-driven flow.

Physical Interpretation

The pressure-gradient force arises from spatial variations in pressure within a fluid, acting on the surfaces of a fluid parcel to produce a net force directed from regions of higher pressure toward lower pressure. Consider a small, imaginary cubic parcel of fluid; the pressure exerts a force perpendicular to each face, proportional to the pressure at that surface times the face area. If pressure is higher on one side of the parcel than the opposite side, the inward push from the high-pressure side exceeds that from the low-pressure side, resulting in a net unbalanced force that accelerates the parcel toward the lower pressure region. In simple physical interpretations, this mechanism often assumes the is incompressible, meaning the parcel's remains constant and volume does not change under variations, allowing focus on the direct effects of pressure differences without complications from compression or expansion. This assumption simplifies the analysis for many geophysical fluids like , where changes are small compared to gradients. The magnitude of the acceleration due to this force is inversely proportional to the fluid's density, such that for a given pressure gradient, denser fluids experience a weaker net acceleration per unit mass, as the same pressure difference must act on more mass. For example, in air with density around 1.2 kg/m³, a pressure gradient of 800 Pa over 300 km yields an acceleration of about 2.2 × 10^{-3} m/s². This inverse density dependence highlights why the force's effect is more pronounced in lighter fluids like the atmosphere than in denser ones like the ocean. Intuitively, this can be visualized like a on , where differences in across the —higher on the windward side—create a net push, propelling the boat forward; similarly, imbalances across the fluid parcel's faces drive its motion through the surrounding medium.

Mathematical Formalism

Derivation from Momentum Equation

The Navier-Stokes equations describe the motion of viscous fluids and serve as the fundamental momentum equations in . In a common form often used for incompressible or low-Mach-number compressible flows, they are expressed as \rho \left( \frac{\partial \mathbf{u}}{\partial t} + (\mathbf{u} \cdot \nabla) \mathbf{u} \right) = -\nabla P + \mu \nabla^2 \mathbf{u} + \rho \mathbf{g}, where \rho is the fluid density, \mathbf{u} is the velocity vector, P is the pressure, \mu is the dynamic viscosity, and \mathbf{g} is the gravitational acceleration vector. This equation represents Newton's second law applied to a fluid element, with the left side denoting the rate of change of momentum per unit volume and the right side accounting for the forces acting on it. The term -\nabla P arises from the stress tensor in the , specifically the isotropic contribution, which integrates over the surface of a to yield a per unit volume directed opposite to the . To isolate the , consider an (setting \mu = 0) without body forces (setting \mathbf{g} = 0) for simplicity; the simplifies to the Euler \rho \frac{D\mathbf{u}}{Dt} = -\nabla P, where \frac{D}{Dt} is the . Dividing through by \rho gives the due to as -\frac{1}{\rho} \nabla P, identifying the per unit mass as -\frac{1}{\rho} \nabla P. This term represents the from differences across the element, driving motion from high to low regions. For a one-dimensional case along the vertical z-direction, consider a fluid parcel of cross-sectional area A and height \Delta z. The net pressure force on the parcel is (P(z) - P(z + \Delta z)) A \approx -\frac{\partial P}{\partial z} \Delta z \, A. The mass of the parcel is \rho A \Delta z, so the force per unit mass is -\frac{1}{\rho} \frac{\partial P}{\partial z}. In hydrostatic equilibrium (where acceleration is zero and viscosity is negligible), this balances gravity: $0 = -\frac{\partial P}{\partial z} + \rho g_z, yielding \frac{\partial P}{\partial z} = -\rho g_z and thus the pressure-gradient force per unit mass -\frac{1}{\rho} \frac{\partial P}{\partial z} = g_z. In dynamic contexts, such as non-equilibrium flow, the same term \frac{\partial P}{\partial z} contributes to the vertical momentum balance alongside inertial and other forces. Thus, the explicit pressure-gradient force term in the momentum equation is -\frac{1}{\rho} \nabla P per unit mass (or -\nabla P per unit volume), encapsulating the effect of spatial variations on .

Vector and Component Forms

The , often denoted as \vec{F}_{pg}, represents the force per unit mass exerted on a element due to spatial variations in , and in its standard form is given by \vec{F}_{pg} = -\frac{1}{\rho} \nabla P, where \rho is the and \nabla P is the . This formulation arises in the momentum equation of , directing the force from regions of higher to lower , perpendicular to isobaric surfaces. The gradient operator \nabla in is defined as \nabla P = \left( \frac{\partial P}{\partial x}, \frac{\partial P}{\partial y}, \frac{\partial P}{\partial z} \right) in Cartesian coordinates, capturing the directional rate of change of along each . Thus, the form encapsulates the collective influence of these partial , with the negative sign ensuring the force opposes the pressure increase. For practical computations, the form decomposes into scalar components in Cartesian coordinates: F_x = -\frac{1}{\rho} \frac{\partial P}{\partial x}, \quad F_y = -\frac{1}{\rho} \frac{\partial P}{\partial y}, \quad F_z = -\frac{1}{\rho} \frac{\partial P}{\partial z}. These components allow for straightforward evaluation in numerical models or analytical solutions, where each term quantifies the force contribution along the respective direction. In cases of variable density, such as baroclinic flows where density gradients exist independently of pressure, an alternative notation expresses the force as -\nabla \left( \frac{P}{\rho} \right), but this holds strictly under the assumption of constant density; otherwise, it introduces an additional term \frac{P}{\rho^2} \nabla \rho to account for density variations. This variation highlights the conservative nature of the force in barotropic conditions (constant density) versus the torque-generating potential in baroclinic atmospheres or oceans. To illustrate, consider a hypothetical pressure field P(x, y, z) = 10^5 + 100x - 50y in a of constant \rho = 1.2 kg/m³ (units simplified for clarity). The is \nabla P = (100, -50, 0) /m, yielding \vec{F}_{pg} = -\frac{1}{1.2} (100, -50, 0) \approx (-83.3, 41.7, 0) m/s². Here, the force magnitude is \sqrt{(-83.3)^2 + (41.7)^2} \approx 93.2 m/s², directed primarily in the negative x-direction toward decreasing , demonstrating how the points downslope along the .

Applications in Fluids

Atmospheric Dynamics

In atmospheric dynamics, the pressure-gradient force (PGF) plays a central role in generating horizontal winds by driving air from regions of to low pressure, often achieving balance with the in the geostrophic approximation for large-scale flows. This balance results in geostrophic winds that flow parallel to isobars, with the wind speed proportional to the PGF magnitude, as the deflects the motion to the right in the (or left in the Southern). Such equilibrium is a of mid-latitude systems, where friction is negligible, allowing the PGF to sustain steady circulations over synoptic scales. Vertically, the PGF's component balances gravitational force in hydrostatic equilibrium, maintaining the atmosphere's layered structure where pressure decreases with altitude, supporting the weight of overlying air without significant vertical acceleration. Deviations from this equilibrium, such as those induced by heating or cooling, create imbalances that accelerate air parcels upward or downward, driving convective motions like those in thunderstorms. For instance, positive from warmer air reduces , enhancing the upward PGF relative to and promoting ascent. These perturbations are crucial for vertical transport in phenomena. A key example of PGF's influence is in the formation of cyclones and anticyclones, where horizontal pressure gradients over low- and high-pressure systems induce rotational flows. In cyclones, steep gradients around the low-pressure core drive inward spiraling winds, intensified by the Coriolis effect, leading to counterclockwise rotation in the and convergent ascent. Conversely, anticyclones feature outward-diverging winds from high-pressure centers, resulting in clockwise rotation and , with gradient strength determining system intensity. The application of PGF varies with scale: on synoptic scales (thousands of kilometers), it predominantly balances the in geostrophic flow, shaping broad patterns like extratropical cyclones. On mesoscales (tens to hundreds of kilometers), ageostrophic effects dominate, allowing PGF to accelerate flows more directly, as in sea breezes or squall lines, where Coriolis influence weakens. In climate models, PGF contributes to global circulation cells, such as the , by driving equatorward surface winds in response to thermal contrasts, with meridional momentum balances involving PGF flux convergence sustaining the cell's upper branch. This dynamical role helps simulate poleward energy transport and ITCZ positioning in general circulation models.

Oceanic Circulation

In oceanic circulation, the pressure-gradient force (PGF) drives motion through distinct barotropic and baroclinic components, reflecting the medium's nature and strong from and variations. The barotropic PGF originates from slopes in sea surface height, producing a depth-independent force that propels basin-scale flows uniformly across the . Conversely, the baroclinic PGF arises from non-parallel isobaric and isopycnal surfaces due to gradients, creating depth-varying forces that induce vertical in currents. These components differ from atmospheric by emphasizing oceanic incompressibility and persistent , which amplify baroclinic effects over large depths. The PGF integrates into geostrophic balance within major ocean gyres, where it equilibrates with the to sustain steady, large-scale circulations along isobars. In subtropical gyres, for instance, subtle sea surface height differences—on the order of 1 meter across thousands of kilometers—generate weak but persistent PGFs that, when balanced by Coriolis deflection, form clockwise () or counterclockwise () patterns. This balance dominates away from boundaries, with altimetry revealing dynamic height anomalies as proxies for the underlying PGF structure. Atmospheric forcing initiates these setups by piling water against coasts, but the geostrophic adjustment propagates the PGF basin-wide. The exemplifies PGF-driven oceanic flow, as its intense northward transport results from barotropic gradients established by wind-induced sea surface elevations and baroclinic gradients from sharp contrasts across the current's front. differences, arising from warmer subtropical waters meeting cooler northern inflows, slope isopycnals and enhance the horizontal PGF, accelerating the current to speeds exceeding 2 m/s while contributing to the . forces from these gradients sustain the stream's separation from the coast, distinguishing it from purely wind-driven components. Vertical aspects of the PGF influence and by linking density-driven horizontal gradients to cross-shore transports that induce or . Sloped isopycnals, tilted by initial or forcings, generate baroclinic PGFs that drive offshore surface flows during upwelling-favorable conditions, drawing nutrient-rich deep water upward to restore geostrophic balance. In downwelling scenarios, onshore steepens isopycnals, intensifying the PGF to promote and water column stabilization. Density-driven PGFs underpin , powering the Atlantic Meridional Overturning Circulation (AMOC) through baroclinic forces from polar sinking of dense water and equatorial . As of 2025, projections indicate AMOC weakening by 18-43% by 2100 in moderate scenarios, with some high-emission models suggesting up to 80% weakening or shutdown after 2100; while abrupt remains debated, Southern Ocean provides some resilience by compensating for reduced overturning. Recent 2025 analyses highlight risks of AMOC slowdown disrupting North Atlantic heat transport, potentially cooling by 1-3°C and altering patterns, though exact impacts remain uncertain. This slowdown disrupts heat transport, cooling the North Atlantic while enhancing Pacific influences on global circulation.

Engineering Contexts

In engineered fluid systems, the pressure-gradient force plays a central role in driving through , where it balances frictional losses to maintain steady of liquids or gases. In , the force arises from the spatial variation in along the length, propelling the against viscous and turbulent resistances. For laminar and turbulent regimes, this is quantified through the Darcy-Weisbach equation, which relates the pressure drop \Delta p to the f, length L, D_h, density \rho, and v: \Delta p = f \frac{L}{D_h} \frac{\rho v^2}{2} This pressure drop, or gradient \frac{\Delta p}{L}, directly drives the flow rate, with laminar flows (low Reynolds numbers) exhibiting linear dependence on the gradient, while turbulent flows show quadratic velocity scaling. Engineers design pipelines to optimize this gradient for efficient transport, such as in water distribution networks where excessive drops increase pumping costs. In applications, pressure gradients within jet engines and rocket nozzles convert high chamber pressures into high-velocity exhaust, generating via the pressure-gradient force acting on the fluid. In rocket nozzles, the force accelerates from the through the converging-diverging geometry, where the in the divergent section expands the flow supersonically. F incorporates this through the term (P_e - P_a) A_e, where P_e is exit pressure, P_a is , and A_e is exit area, supplementing . For example, in engines, stages create initial pressure rises that the gradient exploits for efficient , achieving specific impulses around 300-400 seconds in modern designs. Heating, ventilation, and air conditioning (HVAC) systems rely on controlled gradients to induce in ducts, distributing conditioned air efficiently. Fans generate differences that overcome frictional losses, modeled similarly via the Darcy-Weisbach equation adapted for rectangular ducts using . In typical low-velocity systems (e.g., 5-10 m/s), the gradient ensures uniform flow, with total loss including and dynamic components decreasing downstream. Standards like those from the and Contractors' National Association (SMACNA) specify rates of 0.08-0.1 inches per 100 feet to balance energy use and comfort, preventing uneven distribution in buildings. Contemporary extends pressure-gradient force applications to microfluidic devices, where micro-scale channels (10-100 μm) amplify the force's dominance over other effects like , enabling precise biotech manipulations. Pressure-driven , often via pumps, creates gradients up to 10^4 / to propel or cells through for applications like models and pediatric diagnostics. In 2025 biotech, this facilitates real-time analysis from microliter samples, mimicking physiological for drug testing in conditions such as , with rates controlled to 1-100 μL/min for minimal invasiveness. These devices outperform macro-scale systems by reducing reagent use by orders of magnitude while enhancing resolution. Safety considerations in engineered systems underscore the risks of unmanaged pressure gradients, particularly in and pipelines where excessive values can precipitate structural failures. In embankment , high hydraulic gradients (often exceeding 1) at the downstream toe drive seepage forces that initiate , transporting soil particles and potentially breaching the structure if unmitigated by filters or drains. Historical analyses of large show internal accounts for approximately 47% of failures (1800-1986 data), though overall U.S. indicate around 20-30%. In pipelines, abrupt pressure s—manifesting as steep gradients from closures—induce , generating transient pressures up to 10 times steady-state values that rupture welds or burst pipes, as seen in incidents releasing millions of cubic feet of gas. involves arrestors and gradual operations to limit gradients below material thresholds.

Hydrostatic Balance

In hydrostatic balance, the pressure-gradient force acts vertically to counteract the gravitational force on parcels in a static , resulting in no net motion. This equilibrium is described by the hydrostatic equation, which states that the of pressure equals the negative product of density and the gravitational vector: \nabla P = -\rho \mathbf{g}. Here, the vertical component of the pressure-gradient force precisely balances the weight of the above a given point, preventing vertical . In the atmosphere, where air density decreases with altitude due to compressibility and the , this balance leads to an decrease in with . The characteristic scale over which drops by a factor of e is the atmospheric H, given by H = \frac{RT}{[g](/page/G)}, where R is the specific for dry air (approximately 287 J/kg·K), T is the temperature in , and g is the (about 9.8 m/s²). For typical tropospheric temperatures around 250 K, H is roughly 7-8 km, explaining the rapid falloff observed in Earth's atmosphere. In the , where is nearly constant with depth (varying by only about 2-3% from surface to due to and effects), the hydrostatic balance produces a linear increase in with depth. The at depth h is approximately P = \rho g h, with \rho around 1025 /m³, leading to an increase of about 1 (≈10^5 ) every 10 meters—far steeper than in the atmosphere due to higher . When this balance is disturbed—such as by heating, cooling, or external forcing—the vertical no longer equals the gravitational term, generating buoyancy-driven flows that initiate dynamic motion in the fluid. Modern models increasingly incorporate non-hydrostatic formulations to resolve these deviations, allowing explicit simulation of vertical accelerations in convective processes that the traditional hydrostatic approximation overlooks, as implemented in systems like the ECMWF Integrated Forecasting System.

Distinction from Other Forces

The pressure-gradient force (PGF) differs fundamentally from , which acts as a uniform per unit mass directed downward throughout the , independent of spatial position. In contrast, the PGF is a surface arising from spatial variations in , directed from regions of to low , and its magnitude depends on the \nabla P. This distinction is evident in the hydrostatic balance, where the vertical component of the PGF exactly opposes to maintain equilibrium in a resting , but horizontal PGF components drive motion when unbalanced. Unlike viscous forces, which are dissipative and arise from internal within the , leading to stresses that oppose relative motion between layers, the PGF is an inviscid, conservative derived solely from differences without energy . Viscous forces scale with gradients and , becoming negligible in high-Reynolds-number flows, whereas the PGF remains a primary driver in such regimes. In the Navier-Stokes equations, the PGF -\frac{1}{\rho} \nabla P appears separately from the viscous \nu \nabla^2 \mathbf{u}, highlighting their orthogonal roles in transport. The PGF also contrasts with the , an apparent or resulting from the in a , which deflects moving parcels perpendicular to their velocity without altering speed. The PGF is a real physical based on molecular collisions creating imbalances, always present regardless of fluid motion, while the Coriolis force vanishes for stationary parcels and depends on velocity magnitude. In rotating fluids, such as the atmosphere, these forces interact in balances like geostrophy, where the Coriolis force counters the horizontal PGF to produce straight-line flow parallel to isobars. In the full momentum balance of , the PGF often dominates in large-scale, s where viscous effects are small and rotational influences like Coriolis lead to approximate equilibria, such as in synoptic-scale atmospheric circulations. This dominance stems from the PGF's role in accelerating fluid parcels over broad spatial scales, as captured in the Euler equations for , where it is the primary non-conservative term alongside body forces. A common misconception equates the PGF with ; however, represents a specific net upward force on a displaced parcel due to density-stratified pressure gradients in a , effectively modulating rather than being a general pressure-driven acceleration.

References

  1. [1]
    The Pressure Gradient Force (PGF) – Physics Across Oceanography
    The PGF has the opposite sign to the pressure gradient. It is directed from high to low values of pressure, ie, it is a down-gradient vector.Missing: authoritative sources
  2. [2]
    [PDF] Forces Pressure gradient force (PGF)
    A fluid parcel placed in a pressure gradient is subjected to a net force associated with pressure difference between one side and the other.<|control11|><|separator|>
  3. [3]
    Origin of Wind | National Oceanic and Atmospheric Administration
    Apr 3, 2023 · The pressure gradient force (Pgf) is a force that tries to equalize pressure differences. This is the force that causes high pressure to push ...
  4. [4]
    [PDF] Pressure Gradients
    Page 8. Pressure gradient force ! Tendency for fluids to flow from high pressure to low pressure. Ahrens: Fig. 8.17. Page 9. Horizontal pressure gradient force.Missing: authoritative sources
  5. [5]
    [PDF] Fundamental Fluid Forces - Atmospheric and Environmental Sciences
    We note that the pressure gradient force. (PGF) is proportional to the gradient of the pressure field, and not the pressure itself! P z = 1 p z.
  6. [6]
    [PDF] Lecture 8. Pressure Gradient Force
    • Pressure Gradient Force is expressed as force per unit volume, directed from high to low pressure: PGF = On a contour map of pressure at a given altitude ...
  7. [7]
    [PDF] Chapter 10 Atmospheric Forces & Winds
    Pressure gradient force: A force that acts on air due to spatial differences in pressure. In response to the change in surface pressure a pressure gradient ...Missing: definition | Show results with:definition<|control11|><|separator|>
  8. [8]
    Balloon Rockets - SERC (Carleton)
    In this activity, students will learn that compressed air creates pressure to propel an air balloon. ... Students set up a balloon-rocket system and find ...
  9. [9]
    Physics
    Feb 3, 2003 · These pressure gradients, which can be thought of as a parcel of water being pushed harder from one side than the other cause movement to occur.Missing: explanation | Show results with:explanation<|control11|><|separator|>
  10. [10]
    [PDF] Chapter 6 The equations of fluid motion - Weather in a Tank
    The curved surface provides a pressure gradient force directed inwards that is balanced by an outward centrifugal force due to the anticlockwise circulation of ...<|control11|><|separator|>
  11. [11]
    10.2 What are the important real forces? | METEO 300 - Dutton Institute
    But because the pressure gradient force is minus 1 over the density times the pressure gradient force-- really the pressure gradient acceleration-- is negative.
  12. [12]
    Understanding The Forces Governing Wind
    Oct 21, 2014 · The speed and direction of the wind is governed by three forces; the pressure gradient force (PGF), the Coriolis Force and friction.
  13. [13]
    [PDF] 3 The Navier-Stokes Equation - DAMTP
    ⇢ Du Dt= rP + µr2u In what follows, we'll often divide by the density to write the Navier-Stokes equation as Du Dt = 1 ⇢rP+ ⌫r2u where, as defined earlier, ⌫ = ...
  14. [14]
    [PDF] Atmospheric and Oceanic Fluid Dynamics
    ... Gradient Wind Balance. 94. 2.10. Static Instability and the Parcel Method. 97 ... Pressure coordinates, Section 2.6 (optional, mainly for meteorologists) ...<|separator|>
  15. [15]
    [PDF] Momentum Equation
    The volume of the box cannot change if the parcel is incompressible. Horizontal convergence in the x-direction at the initial time means that the box must.
  16. [16]
    Revisiting an Old Concept: The Gradient Wind* - AMS Journals
    The geostrophic wind arises from a balance of the Coriolis and pressure gradient forces. The gradient wind is then introduced as a refinement of the geostrophic ...
  17. [17]
    Estimation of Winds from GPS Radio Occultations
    The zeroth-order winds that result from the dynamical balance of the pressure gradient and the Coriolis force are the geostrophic winds, obtained by the ...
  18. [18]
    ATMO 203 Tutorials - The Vertical Pressure Gradient Force
    The Vertical Pressure Gradient Force · The force that counteracts gravity and keeps the air up in the atmosphere is called the vertical pressure gradient force.
  19. [19]
    The Ups and Downs of Air Parcels | METEO 3 - Dutton Institute
    The vertical decrease in pressure naturally means that an upward pointing pressure-gradient force exists. In fact, the upward pointing pressure-gradient force ...
  20. [20]
    Investigation of Hydrostatic Imbalance with Field Observations in
    The hydrostatic imbalance of the mean atmospheric state for the acceleration of vertical motions in the vertical momentum balance is investigated using ...
  21. [21]
    The Vertical Profile of Wind and Temperature in Cyclones and ...
    When cyclones are deeper, they may be overall more intense, so that the magnitude of the surface pressure gradient is stronger, the surface winds stronger, and ...
  22. [22]
    Seasonal Aspects of an Objective Climatology of Anticyclones ...
    Anticyclones are governed by different atmospheric dynamics according to their scale. ... pressure or pressure gradient. Thus, the Laplacian of pressure is ...
  23. [23]
    [PDF] Introduction to Mesoscale Meteorology: Part II - twister.ou.edu
    Feb 24, 2015 · The pressure gradient force and Coriolis force are in rough balance. What kind of flow do you get in this case? quasi-geostrophic flow ...
  24. [24]
    The Detailed Dynamics of the Hadley Cell. Part II - AMS Journals
    In this model the HCs exist to reduce the midtropospheric temperature gradient ... models of the basic state for instability and climate studies. Part II ...
  25. [25]
    Role of Continents in Driving the Hadley Cells in - AMS Journals
    Abstract. GCM simulations are used to investigate how forcing that originates over land surfaces influences the Hadley circulation.
  26. [26]
    OC2910 - Module 1: Definitions - Oceanography Department
    The baroclinic pressure gradient is that part of the pressure gradient that is due to the fact that surfaces of constant density are not everywhere parallel to ...Missing: form | Show results with:form
  27. [27]
    Geostrophic balance – Physics Across Oceanography
    Geostrophic balance is the steady-state force balance between the Coriolis and pressure gradient forces, causing flow along isobars.
  28. [28]
    Gulf Stream Dynamics along the Southeastern U.S. Seaboard in
    Mar 1, 2015 · The form drag is the force that results from pressure differences across an obstacle in the flow. It should not be confused with the bottom ...
  29. [29]
    Energetics of Eddy–Mean Flow Interactions in the Gulf Stream ...
    Apr 1, 2015 · A detailed energetics analysis of the Gulf Stream (GS) and associated eddies is performed using a high-resolution multidecadal regional ocean model simulation.
  30. [30]
    Continued Atlantic overturning circulation even under climate ...
    Feb 26, 2025 · Here we show that the AMOC is resilient to extreme greenhouse gas and North Atlantic freshwater forcings across 34 climate models.
  31. [31]
    Darcy-Weisbach Equation: Flow Resistance & Pressure Loss ...
    The Darcy-Weisbach equation can be used to calculate the major pressure and head loss due to friction in ducts, pipes or tubes.
  32. [32]
    [PDF] BASIC PRINCIPLES OF FLOW OF LIQUID AND PARTICLES IN A ...
    equation relates the driving force generated by the pressure gradient over the pipe ... This equation is known as the Darcy-Weisbach equation for the ...
  33. [33]
    Rocket Thrust Equations
    The amount of thrust produced by the rocket depends on the mass flow rate through the engine, the exit velocity of the exhaust, and the pressure at the nozzle ...
  34. [34]
    [PDF] SMACNA Duct Design Fundamentals - Utah ASHRAE
    Pressure Losses – Friction​​ The left‐hand side of the Darcy‐Weisbach Equation, which is the Darcy Equation, calculates the friction loss.
  35. [35]
    Air Flow, Air Systems, Pressure, and Fan Performance - CaptiveAire
    The Darcy-Weisbach Equation should be used for “non-standard” duct type such as flex duct. It is convenient to calculate pressures in ducts using as a base an ...
  36. [36]
    Advancements and Future Perspectives of Microfluidic Technology ...
    Sep 14, 2025 · Similar to continuous microfluidics, it utilizes precisely controlled pumps to generate pressure gradients, driving fluids containing ...
  37. [37]
    Development of microfluidic devices for biomedical and clinical ...
    Aug 17, 2010 · Pressure driven flow allows fluid to be moved via positive or negative displacement, such as syringe or peristaltic pumps. The static nature of ...
  38. [38]
    How Microfluidic Devices are Benefitting Biopharmaceutical ...
    Apr 17, 2024 · Within microfluidic devices, fluid flow can be pressure driven or based on electroosmotic flow. Construction materials include both rigid and ...
  39. [39]
    Piping failures in small dams | Association of State Dam Safety
    Because of this increased hydraulic gradient, the erosion accelerates until the dam breaches. Early dam designs (up until the mid-70s) considered concrete ...Missing: pipelines | Show results with:pipelines
  40. [40]
    Internal Erosion and Piping Evolution in Earth Dams Using an ...
    Mar 21, 2019 · Internal erosion, also known as piping, is one of the major causes of earth dam failures. Piping occurs when flowing water transports soil ...
  41. [41]
    What is Water Hammer? All That You Need to Know! - DFT® Valves
    Water hammer is the result of a pressure surge, or high-pressure shockwave that propagates through a piping system when a fluid in motion is forced to change ...
  42. [42]
    [PDF] Transient and Surge Related Pipe Bursts, Water Loss and Damage ...
    This term is given to the pressure fluctuations that can develop in a pipe when the flow of liquid is suddenly changed. The most common example of water hammer ...
  43. [43]
    [PDF] The Hydrostatic equation. Consider a rectangle of air with ...
    Example: How does the pressure at a fixed elevation change, if the atmosphere warms up? If the atmosphere warms up, the scale height increases, H = RT g. , ...
  44. [44]
    OC2910: Physical Oceanography Basic Concepts
    If the density were constant with depth, the hydrostatic pressure, p, in the ocean would simply be given by the product of the gravitational acceleration, g ...
  45. [45]
    [PDF] Non-hydrostatic modelling at ECMWF
    The current version of the non-hydrostatic IFS model takes approximately three times longer at T3999 resolution (2.5 hours for a 24-hour forecast) due to the ...
  46. [46]
    Introduction to Pressure in Fluid Mechanics
    Pressure is a stress (force per unit area), pressure must be multiplied by the area of the face (dydz) to get dimensions of force.
  47. [47]
    10.8 A Closer Look at the Four Force Balances | METEO 300
    Let the Coriolis force per unit mass be designated as Co=−fV and the pressure gradient force per unit mass as P=−∂Φ∂n . Then the force balances are shown in the ...Missing: formula | Show results with:formula
  48. [48]
    Pressure and buoyancy - Physics
    Nov 12, 1999 · In both cases, the difference in pressure results in a net upward force (the buoyant force) on the object. Example. A basketball floats in a ...Missing: mechanics | Show results with:mechanics