Fact-checked by Grok 2 weeks ago

Richardson number

The Richardson number (Ri) is a dimensionless parameter in fluid mechanics and geophysical flows that quantifies the relative importance of buoyancy forces, arising from density stratification, to the shear forces generated by velocity gradients in a fluid. It serves as a key indicator of flow stability, with values greater than approximately 0.25 typically signifying stable conditions where buoyancy suppresses certain instabilities like Kelvin-Helmholtz, potentially leading to laminar flow, while lower values suggest dynamic instability and enhanced mixing. Named after the British mathematician, physicist, and meteorologist Lewis Fry Richardson (1881–1953), who developed foundational concepts in numerical weather prediction and atmospheric stability, the number is expressed in its gradient form as Ri = \frac{g}{\theta_v} \frac{\partial \theta_v / \partial z}{(\partial U / \partial z)^2 + (\partial V / \partial z)^2}, where g is gravitational acceleration, \theta_v is virtual potential temperature, z is the vertical coordinate, and U and V are horizontal velocity components. Several variants of the Richardson number exist to suit different observational and modeling contexts, including the flux Richardson number (Rf), which compares buoyant production of turbulent kinetic energy to its mechanical production, and the bulk Richardson number (Rb), which approximates gradients using finite differences over a layer thickness. The flux form is Rf = -\frac{g / \theta_v \langle w' \theta_v' \rangle }{ \langle u' w' \partial U / \partial z + v' w' \partial V / \partial z \rangle } (with conventions such that Rf > 0 indicates stable stratification), with a theoretical upper limit of 1 when buoyancy destruction equals shear production; observed values rarely exceed 0.2–0.3 due to incomplete suppression of shear production. The bulk form, Rb = \frac{g \Delta \theta_v \Delta z / \theta_v}{(\Delta U)^2 + (\Delta V)^2}, is particularly useful for coarse-resolution data in atmospheric or oceanic profiles. These forms highlight the number's role in diagnosing transitions between laminar and turbulent regimes, informed by theoretical limits like the Miles-Howard theorem, which posits a critical gradient Ri of 0.25 for inviscid stability. In practice, the Richardson number is applied across disciplines to model stratified in the atmosphere, oceans, and engineering systems, such as predicting stability for or mixing in reservoirs and estuaries. In contexts, for instance, a bulk Ri below 1 facilitates strong turbulent mixing across interfaces, while values above 1 promote layered flows with reduced dissipation; experimental studies show mixing peaking near Ri ≈ 1. Similarly, in meteorological forecasting, low Ri values signal potential for , aiding hazard mitigation. Its dimensionless nature allows universal scaling across systems, from laboratory grid-stirred tanks to planetary layers, underscoring its enduring utility in understanding buoyancy-shear interactions.

Fundamentals

Definition

The Richardson number is named after (1881–1953), a British mathematician, physicist, and meteorologist renowned for his pioneering contributions to . During , while serving as an ambulance driver on the Western Front, Richardson began developing systematic methods to forecast weather using mathematical computations, including early concepts for evaluating atmospheric stability. He first conceptualized the parameter in his 1920 paper "The supply of energy from and to atmospheric eddies," where it emerged as a criterion for assessing energy transfer and turbulence in stratified atmospheres. This foundational work laid the groundwork for Richardson's influential 1922 book Weather Prediction by Numerical Process, which detailed algorithms for numerical forecasting and highlighted the parameter's role in understanding atmospheric dynamics. Originating in early 20th-century amid efforts to model complex weather patterns manually, the Richardson number has since evolved into a cornerstone of research. At its core, the Richardson number serves to quantify the relative strength of forces—driven by variations in a stratified —compared to forces generated by velocity gradients across the flow. This balance determines whether stabilizing effects dominate or destabilizing promotes mixing in layered fluids. In general, the aids in predicting shifts between laminar and turbulent flow states or between stable and unstable configurations in mediums like air and , extending its utility from atmospheric studies to diverse geophysical and engineering contexts.

Formulation and Variants

The gradient Richardson number, often denoted as Ri_g, quantifies the ratio of buoyancy effects due to vertical density stratification to shear production of turbulence in a fluid flow. It is defined as Ri_g = \frac{ \frac{g}{\rho_0} \left( -\frac{\partial \rho}{\partial z} \right) }{ \left( \frac{\partial u}{\partial z} \right)^2 }, where g is gravitational acceleration, \rho_0 is a reference density, \partial \rho / \partial z < 0 indicates stable stratification with density decreasing upward, u is the horizontal velocity, and z is the vertical coordinate. This formulation arises from dimensional analysis comparing the potential energy associated with buoyancy restoration (proportional to g (-\partial \rho / \partial z) / \rho_0) to the kinetic energy of vertical shear (proportional to (\partial u / \partial z)^2), as derived in the context of turbulence kinetic energy (TKE) balance where buoyancy acts to suppress mixing while shear promotes it. In atmospheric applications, it is equivalently expressed using virtual potential temperature \theta_v as Ri_g = \frac{ \frac{g}{\theta_v} \frac{\partial \theta_v}{\partial z} }{ \left( \frac{\partial U}{\partial z} \right)^2 + \left( \frac{\partial V}{\partial z} \right)^2 }, where U and V are horizontal wind components, assuming hydrostatic balance and Boussinesq incompressibility. The densimetric Richardson number, Ri_d, is the density-based gradient form under the Boussinesq approximation for flows with small density variations treated as buoyancy perturbations. It is given by Ri_d = \frac{ N^2 }{ \left( \frac{\partial u}{\partial z} \right)^2 }, where N^2 = -\frac{g}{\rho_0} \frac{\partial \rho}{\partial z} is the squared buoyancy frequency, with \partial \rho / \partial z < 0 for stable stratification. This form is equivalent to the gradient Richardson number and is commonly used in oceanographic and estuarine flows where density gradients from salinity or temperature drive stratification, such as in lock-exchange experiments. The flux Richardson number, Ri_f, measures the ratio of turbulent buoyancy flux to shear production in the surface layer, particularly useful for parameterizing eddy diffusivities in stably stratified turbulence. It is expressed as Ri_f = \frac{ \frac{g}{T_v} \overline{w' \theta_v'} }{ \overline{u' w'} \cdot \frac{\partial u}{\partial z} }, where T_v is a reference virtual temperature, w' \theta_v' is the vertical turbulent heat flux, and - \overline{u' w'} is the magnitude of the vertical momentum flux (related to wind stress). This variant derives from the TKE equation by taking the ratio of the buoyancy destruction term g \overline{w' \theta_v'} / T_v to the shear production term - \overline{u' w'} (\partial u / \partial z), assuming horizontal homogeneity, negligible subsidence, and that turbulent fluxes represent the dominant transport mechanisms near the surface. It is often related to the gradient form via eddy diffusivity closures, such as Ri_f \approx Ri_g / (1 + Ri_g), but emphasizes measurable fluxes over local gradients. The bulk Richardson number, Ri_b, provides a layer-averaged estimate suitable for finite-depth profiles where local gradients are unavailable, such as in boundary layer modeling. It is formulated as Ri_b = \frac{ \frac{g}{T_{v0}} \Delta \theta_v \Delta z }{ (\Delta U)^2 + (\Delta V)^2 }, where T_{v0} is the surface virtual temperature, \Delta \theta_v and \Delta U, \Delta V are differences in virtual potential temperature and horizontal wind components over height difference \Delta z. This arises as an integral approximation to the gradient form, integrating buoyancy potential over the layer thickness against the squared velocity difference, under assumptions of logarithmic profiles or bulk transfer relations for practical computations like surface layer stability. It is commonly used iteratively with to estimate fluxes from mean profiles. In convective heat transfer, the thermal Richardson number, Ri, assesses the relative importance of natural (buoyancy-driven) versus forced convection. It is defined as Ri = \frac{Gr}{Re^2}, where Gr is the (Gr = g \beta \Delta T L^3 / \nu^2, with \beta the thermal expansion coefficient, \Delta T temperature difference, L characteristic length, and \nu kinematic viscosity) and Re is the (Re = U L / \nu, with U forced flow speed). This ratio derives from balancing the buoyancy force scale (Gr) against inertial forces in forced flow (Re^2), indicating forced convection dominance when Ri \ll 1, natural convection when Ri \gg 1, and mixed regimes in between, under incompressible flow and constant properties. These formulations share common assumptions, including incompressible flow via the Boussinesq approximation (density variations only in buoyancy term), hydrostatic balance (pressure gradients support stratification without vertical acceleration), and stable vertical stratification (\partial \rho / \partial z < 0). The original gradient formulation in atmospheric contexts was pioneered by in his analyses of atmospheric turbulence and diffusion in the early 20th century. Subsequent variants, such as the flux and bulk forms, build on this foundation, while the thermal Richardson number arises in engineering applications of convective heat transfer.

Physical Interpretation

Buoyancy Versus Shear

The buoyancy term in the Richardson number arises from density gradients in a stratified fluid, which store potential energy that resists vertical displacements and thereby suppresses turbulent mixing when stratification is stable. In such conditions, denser fluid layers remain below lighter ones, creating a restoring force that inhibits the breakdown of layered structures into chaotic motion. In contrast, the shear term reflects kinetic energy associated with velocity gradients across the fluid layers, where differences in horizontal flow speeds generate mechanical energy that can drive turbulence through instability mechanisms. High shear levels provide the energy input necessary to overcome stratification, promoting the vertical exchange of momentum and scalars by fostering instabilities that lead to enhanced mixing. The Richardson number encapsulates the competition between these effects as a ratio, where values greater than approximately 0.25 signify buoyancy dominance, resulting in persistent stable layering that limits fluid interchange. Conversely, values less than 0.1 indicate shear dominance, allowing kinetic energy to prevail and facilitate widespread turbulent mixing. This balance determines whether the flow remains orderly or transitions toward disorder, with the number serving as a diagnostic for the relative influence of gravitational versus inertial forces. From an energy viewpoint, the Richardson number represents the ratio of buoyant production or destruction of turbulent kinetic energy to its production by shear, quantifying how stratification either replenishes or depletes turbulence relative to mechanical generation. In simple shear flows, such as those between parallel plates with imposed velocity differences, this ratio highlights how increasing stratification reduces overall turbulent energy levels by converting kinetic energy back into potential energy, thereby damping vertical motions. These interpretations rely on assumptions of horizontal homogeneity in the flow field, ensuring that variations occur primarily in the vertical direction without lateral influences complicating the balance. Additionally, the framework typically excludes other external forcings, such as rotational effects from the Coriolis force, to isolate the - interaction.

Flow Stability

In stratified flows, the Richardson number serves as a key parameter for assessing stability regimes. The Miles-Howard theorem establishes that a sufficient condition for inviscid stability is a local gradient Richardson number exceeding 1/4 everywhere in the flow, preventing the growth of infinitesimal disturbances. Flows with Ri > 0.25 are dynamically stable, as forces overwhelm production of , leading to laminar-like conditions. In contrast, Ri < 0 indicates inherent instability, arising from adverse density gradients that promote convective overturning and potential energy release through vertical displacements. The transition between stability and turbulence is closely tied to Richardson number thresholds. When Ri < 0, convective instability dominates, driving vigorous mixing independent of shear. For 0 < Ri < 1/4, shear-driven instabilities emerge, notably , which roll up at the density interface and cascade into turbulence if energy extraction from the mean flow exceeds dissipation. These regimes highlight how low Ri values enable the conversion of kinetic energy into turbulent motions, while higher values suppress them. Laboratory experiments have validated these criteria in controlled stratified flows. Wind tunnel and tank studies using miscible fluids with imposed shear and density gradients confirm that instabilities onset near Ri ≈ 0.25, with billow formation and turbulent breakdown occurring below this threshold, aligning with theoretical predictions. Such validations underscore the robustness of Ri-based diagnostics across varying Reynolds numbers. Theoretical frameworks further bound stability through turbulence modeling. Prandtl's mixing length theory links the Richardson number to reductions in eddy viscosity, positing that stable stratification limits the scale of turbulent eddies, thereby diminishing momentum and scalar transport as Ri increases. This suppression mechanism explains the transition to subdued mixing in high-Ri flows, complementing the Miles-Howard criterion by addressing finite-amplitude effects.

Applications in Geophysics

Meteorology

In meteorology, the Richardson number serves as a critical parameter for assessing atmospheric stability and turbulence, particularly through its atmospheric variants tailored to different observational contexts. The gradient Richardson number evaluates vertical profiles of potential temperature and wind shear, often derived from radiosonde data to diagnose stability in the planetary boundary layer and free atmosphere. The flux Richardson number, in contrast, relates buoyant production to mechanical production of turbulence, commonly applied in the surface layer to parameterize heat and momentum fluxes under varying stability conditions. A primary application involves predicting stability in the nocturnal boundary layer, where high gradient Richardson numbers indicate strong stratification that suppresses vertical mixing and turbulence. In stable conditions, as observed at sites like Cabauw, Netherlands, the flux Richardson number remains constant under z-less stratification, reflecting a turbulence-limited state where buoyant destruction dominates mechanical production. Similarly, in frontal zones, low Richardson numbers signal enhanced wind shear relative to buoyancy, often leading to clear air turbulence; diagnostic tendencies of the Richardson number, incorporating frontogenetical effects, have successfully forecasted such events by predicting drops below critical thresholds near upper-level fronts. Observational methods integrate these variants into numerical weather prediction models, such as the European Centre for Medium-Range Weather Forecasts (ECMWF) system, where eddy diffusivities for vertical diffusion are parameterized as functions of the local to represent turbulent mixing in stable layers. In stable conditions above the boundary layer, the eddy-diffusivity mass-flux approach suppresses mixing when the Richardson number exceeds 0.25, aligning with observational profiles from radiosondes. Case studies highlight the Richardson number's role in severe weather dynamics. In thunderstorms, low Richardson numbers in the boundary layer (typically -2 ≤ Ri ≤ 0) near low-level inversions facilitate convective overturning by promoting shear-driven instability, as evidenced in analyses over Kano, Nigeria, where such values correlated with frequent storm occurrences. For fog formation, high modified Richardson numbers (>1) in the near-surface layer indicate dynamic stability that traps moisture under nocturnal inversions, favoring radiation fog persistence by inhibiting turbulent dispersion, as demonstrated in classifications at Christchurch Airport, New Zealand. Post-2020 advancements in high-resolution models have enhanced turbulence forecasting by explicitly resolving Richardson number distributions. For instance, large-eddy simulations at 35 m resolution reproduced events over in 2020, identifying regions with Ri < 0.25 linked to Kelvin-Helmholtz and validating predictions against onboard flight . High-resolution observations, such as those from LITOS balloons in tropopause folds, reveal Richardson number minima (around 0.21–0.32) driving turbulence in the lower , underscoring the need for refined parameterizations in models like ECMWF's Integrated System to better capture these features. More recent studies as of 2025 project increasing trends in upper-atmospheric under , with declining Richardson numbers near jet streams potentially exacerbating over the North Atlantic, as analyzed in high-resolution ERA5 reanalysis.

Oceanography

In oceanography, the Richardson number is formulated as the gradient Richardson number, Ri = \frac{N^2}{\left( \frac{\partial u}{\partial z} \right)^2 }, where N is the buoyancy frequency and \frac{\partial u}{\partial z} is the vertical of the horizontal velocity. The squared buoyancy frequency is given by N^2 = -\frac{[g](/page/Gravitational_acceleration)}{\rho} \frac{\partial \rho}{\partial z}, with g as , \rho as potential , and \frac{\partial \rho}{\partial z} as the vertical gradient. This formulation accounts for density variations driven by both temperature and salinity through the equation of state for seawater, which introduces nonlinear effects unlike the primarily temperature-dominated atmospheric case. The Richardson number plays a key role in evaluating double-diffusive convection within the ocean's , where low Ri values indicate regions prone to shear-enhanced mixing that can disrupt or amplify salt-fingering instabilities. In such stratified layers, Ri < 0.25 often signals the onset of turbulent patches that interact with double-diffusive processes, leading to enhanced diapycnal fluxes of heat and salt despite overall stable . Similarly, in zones, low Ri < 1 promotes vertical mixing that entrains nutrient-rich waters to the surface, sustaining primary productivity in coastal ecosystems like those off . This shear-driven destabilization, observed during wind-forced events, facilitates delivery without requiring complete overturning of the . Low Richardson numbers also underpin instabilities in fronts and internal fields, linking to the generation of internal solitary and Kelvin-Helmholtz billows that drive turbulent dissipation. In fronts, Ri \approx 0.25 marks the threshold for Kelvin-Helmholtz , where overcomes to produce layered , as evidenced by direct observations of and profiles. These processes have been quantified using autonomous gliders and moored instruments, which reveal sporadic Ri < 1 events within propagating internal solitary , contributing to fine-scale mixing in the upper . Such instabilities are particularly prominent in regions like the , where they enhance energy transfer from large-scale flows to dissipation. In ocean general circulation models like the Modular Ocean Model version 6 (MOM6), the Richardson number parameterizes diapycnal to represent unresolved vertical mixing in stratified flows. The Pacanowski-Philander scheme, commonly implemented in MOM6, suppresses as Ri increases above a critical value of 0.7, yielding realistic profiles of turbulent dissipation in the and abyss. This approach ensures that modeled mixing aligns with observations from moored profilers, preventing excessive in stable layers while allowing enhanced transport during low-Ri events. Recent advances since 2015 have integrated satellite altimetry with in situ data to estimate bulk Richardson numbers in mesoscale eddies, providing global maps of potential instability and mixing efficiency. By deriving geostrophic shear from sea-surface height anomalies and pairing it with climatological buoyancy frequencies, studies have shown that eddies often exhibit Ri \sim O(1), fostering submesoscale fronts with elevated diapycnal fluxes. These bulk estimates, validated against glider transects, highlight eddies' role in modulating carbon export through shear-induced turbulence in the subtropical gyres. As of 2025, further developments include quantifying Ri's influence on subsurface turbulent mixing that amplifies central Pacific El Niño–Southern Oscillation (ENSO) variability, and evaluations of Ri-based parameterizations in gray-zone resolutions for improved turbulence representation in coastal and open-ocean models like CROCO.

Applications in Engineering

Aviation

In aviation, the Richardson number serves as a critical diagnostic for assessing atmospheric and predicting hazards that affect flight , particularly in regions of strong vertical and . The bulk Richardson number (Ri_b), computed over a vertical layer, is employed to forecast mountain wave , where values below 1 indicate conditions conducive to sustained shear-driven and the potential formation of rotor clouds beneath wave crests. These rotor clouds, often associated with low Ri_b layers near , signal hazardous low-level for navigating mountainous routes, as observed in numerical models of lee-side wave breaking. The Richardson number (Ri_g), evaluating local at specific altitudes, aids in in-flight detection of layers prone to . During flight, pilots and systems monitor Ri_g to identify regions where suppression is insufficient against forces, often near streams, enabling proactive altitude adjustments to avoid (CAT). This application draws from airborne measurements during 1990s verification flights, where low Ri_g values correlated with turbulence encounters in stable air masses. Safety protocols incorporate Richardson number thresholds indirectly through turbulence forecasting models that inform (FAA) avoidance guidelines, such as maintaining separation from convective tops where low Ri signals enhanced . Analyses of severe incidents from 1990 to 1996, including 44 U.S. air carrier events, linked many to entrance regions with relative Ri minima, emphasizing the need for Ri-based diagnostics in processes for structural integrity under stratified flow conditions. For instance, low Ri environments near s contributed to accidents like the 1990 TRW event over , where deep-layer instability amplified hazards. Real-time estimation of Ri profiles relies on technologies like and wind profilers, which provide vertical wind and temperature data for approach path monitoring at airports. These systems compute Ri_g up to 1.5 km altitude, detecting low-level and risks during landing, as demonstrated in field experiments combining temperature profiles with profiler winds. In plateau airports, -derived Ri gradients have quantified multi-scale shear, supporting safer operations in complex terrain. Current practices integrate Richardson number diagnostics into next-generation via Automatic Dependent Surveillance-Broadcast (ADS-B) data, enabling Ri-informed alerts. As of 2024, ADS-B tracks from global flights facilitate post-event Ri analysis and real-time forecasting enhancements, such as in Chinese airspace where low Ri regions near jet streams were validated against pilot reports, improving alert accuracy in high-traffic corridors.

Thermal Convection

In thermal convection, the Richardson number (Ri) serves as a key dimensionless parameter to characterize the relative importance of buoyancy-driven natural convection versus forced convection in engineering heat transfer systems. Defined as the ratio of the Grashof number (Gr), which quantifies buoyancy forces, to the square of the Reynolds number (Re), which represents inertial forces, Ri is expressed as \text{Ri} = \frac{\text{Gr}}{\text{Re}^2}. This formulation allows engineers to delineate flow regimes: when Ri ≫ 1, natural convection dominates due to strong buoyancy effects; when Ri ≪ 1, forced convection prevails as inertial forces overwhelm buoyancy; and intermediate values (typically 0.1 < Ri < 10) indicate mixed convection where both mechanisms interact significantly. In engineering applications, is instrumental in designing systems like solar collectors and units, where buoyancy-induced must be managed to optimize . For instance, in hot water storage tanks integrated with solar systems, Ri analysis helps predict stability and mixing, using tank height as the scale to model outlet temperatures and prevent unwanted that degrades stratification. Higher Ri values promote effective thermal layering, enhancing energy retention for peak-shifting in residential and commercial setups, while lower values signal excessive mixing that reduces storage performance. This approach, combined with considerations, informs inlet device designs and optimizations in nondimensional models for chilled water or domestic hot-water tanks. For mixed convection scenarios, correlations for the Nusselt number (Nu), which measures convective heat transfer enhancement, are often adjusted based on Ri to account for buoyancy contributions in pipes and enclosures. In rod bundle geometries simulating nuclear fuel assemblies or pipe flows, experimental studies show Nu increasing with local Ri (ranging from 0.25 to 4.3) due to elevated heat flux along the flow path, with a proposed correlation Nu/Nu₀ = 1 + 0.55 ln(1 + Ri²) capturing the transition from forced to mixed regimes beyond standard laminar predictions. Similarly, in lid-driven enclosures, numerical validations reveal that Nu rises with increasing Ri (up to 10), as buoyancy strengthens circulation and heat transfer rates, particularly at moderate Re (e.g., 100) and low Prandtl numbers (e.g., 0.71). These trends highlight Ri's role in refining heat exchanger designs, where decreasing Ri boosts Nu by favoring forced convection. Laboratory studies on mixed convection setups, akin to modified Rayleigh-Bénard configurations with imposed shear, demonstrate Ri's direct influence on . Experiments in vented, differentially heated enclosures identify three regimes based on Ri: for Ri < 10, alters flow structures and yields Nu scalings closer to pure forced limits, while higher Ri enhances vertical heat transport through plumes, increasing overall flux by up to 20-30% compared to non-buoyant cases. These findings, validated against data, underscore Ri's utility in quantifying thresholds like Rayleigh-Taylor effects on convective onset. Recent advancements in (CFD) have integrated Ri for simulating complex thermal convection in HVAC and systems, with post-2020 validations improving predictive accuracy. In ventilated enclosures modeling HVAC heat dissipation (e.g., electronic cooling fins), CFD using Fluent shows Nu correlating as Nu_ave = a Re^b Ri^c (R² = 0.9925) for 0.1 ≤ Ri ≤ 10, where higher Ri elevates thermal performance by 15-25% through dominance, addressing gaps in 3D fin analyses. For passive cooling, guidelines from OECD-NEA emphasize for buoyancy-driven flows at high Gr, with Ri characterizing mixed regimes in thermal shock scenarios; validations like Cutrono et al. (2020) achieve ±3% accuracy in cold leg using SST k-ω models incorporating terms. Hybrid RANS/LES approaches in post-2020 studies further refine Ri-based predictions for reactor pools and small modular designs, reducing uncertainties in by 10-20% via experimental benchmarks.

References

  1. [1]
    2.2: Dimensionless Numbers - Engineering LibreTexts
    Jul 14, 2021 · The Richardson number Ri is named after Lewis Fry Richardson (1881-1953). It is the dimensionless number that expresses the ratio of the ...<|control11|><|separator|>
  2. [2]
    [PDF] Contents 1. The Richardson Number - atmo.arizona.edu
    The Richardson Number (Rf) is the ratio of buoyant to mechanical production, characterizing thermal stability of flow. Rf > 0 is stable, Rf < 0 is unstable, Rf ...
  3. [3]
    [PDF] Laboratory Studies of Turbulent Mixing
    This change is quantified by a flux Richardson number Rfc, defined as rate of change in potential energy divided by power (rate of energy) exerted by the ...
  4. [4]
    The supply of energy from and to atmospheric eddies - Journals
    Richardson Lewis Fry. 1920The supply of energy from and to atmospheric eddiesProc. R. Soc. Lond. A97354–373http://doi.org/10.1098/rspa.1920.0039. Section.
  5. [5]
    100 Years of L. F. Richardson's Weather Prediction by Numerical ...
    Apr 6, 2022 · Richardson, L. F., 1922: Weather Prediction by Numerical Process. Cambridge University Press, xii + 236 pp.
  6. [6]
    Weather Prediction by Numerical Process
    The idea of forecasting the weather by calculation was first dreamt of by Lewis Fry Richardson. He set out in this book a detailed algorithm for systematic ...
  7. [7]
    [PDF] Mixing in a stratified sh ai" flow: energetics and sampling
    (1992) for Richardson numbers from 0 to 1 and for Prandtl numbers between. 0.1 and 2. The results indicate that mixing efficiency. R I varies with turbu- lent ...
  8. [8]
    [PDF] in the planetary boundary layer - DTIC
    Local values of Richardson Number, Ri, Profile contour Number, a, and. Deacon Number, P, at indicated heights from class averages of free-air potential.
  9. [9]
    Estuary classifications based on fluid dynamics principles
    We can derive a single bulk Richardson number by replacing the density and ... The vertical salinity gradient is very small, leading to large densimetric Froude ...Missing: variants thermal
  10. [10]
    [PDF] Overview of density flows and turbidity currents
    ... bulk Richardson number. Ri, which is equal to the reciprocal of the square of the densimetric Froude number. For instance, the following expression was ...Missing: variants | Show results with:variants
  11. [11]
    [PDF] Boundary Layer Meteorology Chapter 5
    the above equation reduces to the more common form of the flux Richardson number: Define the flux Richardson number, Rif : v v f i. i j j g w. Ri u. u u x. ′ ′θ.
  12. [12]
    On the flux Richardson number in stably stratified turbulence
    Jun 8, 2016 · The flux Richardson number is traditionally defined as the ratio of the buoyancy flux B to the production rate of turbulent kinetic energy P .
  13. [13]
    [PDF] Fundamentals of Atmospheric Modeling - Stanford University
    Jan 30, 2002 · Bulk Richardson number. Ratio of buoyancy to mechanical shear. Rib = g θ v zr. ( )− θ v z0,h. ( ). [. ] zr − z0,m. (. )2 θ v z0,h. ( )u zr. ( )2.
  14. [14]
    [PDF] bulk richardson number scheme in aermod
    ... calculated based on the value of θ* calculated from the Bulk Richardson Number Scheme. The equivalent cloud cover is calculated as follows, based on van ...
  15. [15]
    [PDF] Buoyancy effect on the flow pattern and the thermal performance of ...
    Sep 26, 2016 · The Richardson number rep- resents the importance of the natural convection with respect to the forced convection. Usually the range in which ...
  16. [16]
  17. [17]
    [PDF] Nondimensional convection numbers modeling thermally stratified ...
    This paper is focused on establishing the significance of the convection numbers in conjunction with the Reynolds number for modeling the thermal stratification ...
  18. [18]
  19. [19]
  20. [20]
    Buoyancy generated turbulence in stably stratified flow with shear
    Apr 20, 2006 · The energy evolution in buoyancy-generated turbulence subjected to shear depends on the gradient Richardson number and the stratification ...Missing: versus | Show results with:versus
  21. [21]
    Flow measurement in the southern coast of the Caspian Sea
    Thus, for Richardson number greater than 1, buoyancy force (driven by the vertical water density gradient) dominates. In this condition, the stratification ...
  22. [22]
    Critical flux Richardson number for Kolmogorov turbulence enabled ...
    Feb 24, 2019 · In stably stratified flows, the flux Richardson number is a measure of the ratio between buoyancy destruction and shear production of ...
  23. [23]
    Large eddy simulation of stably stratified open channel flow
    Nov 21, 2005 · ... Richardson number is not observed. The vertical Froude number is a ... is less than 0.1 as long as the dimensional channel height is ...
  24. [24]
    [PDF] Environmental Fluid Dynamics: Lecture 19 - Jeremy A. Gibbs
    • This ratio is used to determine which factor ”wins” and. whether the flow becomes turbulent or not. • Examples include the Reynolds number, Richardson number ...
  25. [25]
    [PDF] Note on a paper of John W. Miles
    Miles's theorem and the semicircle theorem limit the values of the Richardson number and the complex wave velocity which are accessible to unstable modes.
  26. [26]
    [PDF] Instability and Turbulence in a Stratified Shear Layer - DTIC
    An important result of their experiment is verification of the Miles-Howard sufficiency condition for stability. For cases when the Richardson number was.
  27. [27]
    Fluid Mixing in Stratified Gravity Currents: The Prandtl Mixing Length
    3 Apr 2009 · A constant eddy-viscosity model provides a poor description of the physics of mixing, but a Prandtl mixing length model relating momentum and ...
  28. [28]
  29. [29]
    The Turbulent Structure of the Stable, Nocturnal Boundary Layer in
    An important application of z-less stratification is that both the Richardson number and flux Richardson number should become constant in the stable boundary ...
  30. [30]
    Application of a Diagnostic Richardson Number Tendency to a Case ...
    An index for the development of clear air turbulence on a synoptic scale is derived using the time rate of change of the Richardson number following the air ...
  31. [31]
    [PDF] of vertical diffusion - ECMWF
    Jan 20, 1994 · NWP model at the ECMWF by eddy diffusivities that depend upon the local. Richardson number (Louis et al. 1981). The diffusivities are chosen ...
  32. [32]
    [PDF] Part IV: Physical Processes - ECMWF
    The Eddy-Diffusivity Mass-Flux (EDMF) approach is used to compute the ... Richardson number becomes superior to 0.25). In stable layers situated above ...
  33. [33]
    Richardson number, vertical wind shear and storm occurrences at ...
    The Richardson number with which storm occurrences are most common is bi-modal in both lower and middle troposphere: −2⩽Ri⩽0 and Ri⩽ −10 in the boundary layer ...Missing: inversions | Show results with:inversions
  34. [34]
    Fog type classification using a modified Richardson number for Christchurch, New Zealand
    ### Summary of Modified Richardson Number for Fog Type Classification
  35. [35]
    (PDF) Clear Air Turbulence Resolved by Numerical Weather ...
    Jun 5, 2023 · In this study, a clear-air turbulence event was reproduced using a high-resolution (250 m) large-eddy simulation in the Weather Research and ...
  36. [36]
    High‐Resolution Observations of Turbulence Distributions Across ...
    Feb 13, 2021 · In this study, we interpret two vertical turbulence measurements. We acquired these uninterrupted high-resolution dissipation rate profiles ...
  37. [37]
    Richardson Number and Ocean Mixing: Towed Chain Observations in
    To examine the importance of salinity to the density gradients, temperature and conductivity data from the collocated temperature–conductivity (T–C) sensors ...
  38. [38]
    Finescale Instabilities of the Double-Diffusive Shear Flow in
    Rigorous instabilities are predicted to occur for Richardson numbers as high as—or even exceeding—unity. The inferences from the linear theory are supported by ...
  39. [39]
    [PDF] Double-Diffusive Convection - WHOI GFD
    At smaller Richardson numbers, horizontal shear interrupts the vertical flux of heat and salt [27]. The main interest in salt fingers within the oceanographic ...
  40. [40]
    A new observationally motivated Richardson number based mixing ...
    Feb 5, 2013 · 2.6 Estimation of Richardson Number. [35] The Richardson number is highly scale-dependent, with the instantaneous value of the Richardson number ...
  41. [41]
    [PDF] Ocean Mixing by Kelvin-Helmholtz Instability
    In this article, we review the history and current status of research into KH instability with a focus on its role in the energy cascade from oceanic internal.Missing: fronts | Show results with:fronts
  42. [42]
    Marginal Instability Within Internal Solitary Waves - AGU Journals
    Apr 29, 2021 · Kelvin-Helmholtz (KH) instability is a ubiquitous process that converts large-scale flows to turbulent mixing in the stratified ocean. This ...Missing: gliders | Show results with:gliders
  43. [43]
    Enhanced Mixing Induced by Near-Inertial Waves Inferred by Glider ...
    Nov 9, 2023 · A diffusivity elevated by one order of magnitude, the occurrence of high near-inertial velocity shears, and the low (≤1) Richardson numbers were ...
  44. [44]
    Microstructure and Mooring Observations of Enhanced Mixing in the ...
    Aug 12, 2020 · The Richardson number (Ri) time series (Figure 11d) estimated from the moored CTD and ADCP data of RM1 indicated that the water column showed ...
  45. [45]
    A Modified Vertical Mixing Parameterization for Its Improved Ocean ...
    Jan 1, 2019 · In this paper, the impacts of two different vertical mixing schemes on the tropical Pacific temperature simulations are investigated using ...
  46. [46]
    [PDF] Simulation of Deep Cycle Turbulence by a Global Ocean General ...
    Abstract Deep cycle turbulence (DCT) is a diurnally oscillating turbulence that penetrates into a stratified shear layer below the surface mixed layer, ...
  47. [47]
    Full‐Depth Global Estimates of Ocean Mesoscale Eddy Mixing From ...
    Sep 9, 2020 · Here we propose a method to estimate eddy mixing using satellite observations and readily available observations of absolute salinity SA, ...Missing: post- | Show results with:post-
  48. [48]
    Eddy Dynamics from Satellite Altimetry - The Oceanography Society
    Oct 2, 2015 · Satellite altimetry observes ocean eddies, providing global maps of their distribution, velocity, and transport of heat and salt.
  49. [49]
    [PDF] Turbulence and mountain waves - ACP
    Oct 2, 2009 · 5 shows the bulk Richardson number, Ri . Values less than. 1 are generally considered necessary to sustain wind-shear driven turbulence, less.
  50. [50]
    Near-Surface Characteristics of the Turbulence Structure during a ...
    Mountain waves and rotors are known to reach particularly striking amplitude and strength there. ... Time series of bulk Richardson number RB in the layer between ...
  51. [51]
    [PDF] AIRBORNE VERIFICATION OF ATMOSPHERIC TURBULENCE ...
    The Richardson number in this study was derived based upon ten flights over the United States in 1995. Airborne turbulence observations were taken by the ...<|separator|>
  52. [52]
    (PDF) Evaluating effectiveness of the FAA's CIT avoidance guidelines
    location of the turbulence encounter. Figure 15: Computed Richardson number (Ri) from the. 12 UTC initialization of the RUC NWP model at the. location of the ...
  53. [53]
    [PDF] Characterizing the Severe Turbulence Environments Associated ...
    The results of these analyses indicate a predominance of severe accident-producing turbulence within the entrance region of a jet stream at the synoptic scale.
  54. [54]
    Determination of Richardson number profile from remote sensing ...
    A number of remote-sensing wind-measuring instruments such as LIDAR and radar wind profilers [6] have been used for monitoring of low-level wind shear and ...
  55. [55]
    Multi-Scale Wind Shear at a Plateau Airport: Insights from Lidar and ...
    The vertical gradient of total horizontal wind speed was used in the Richardson number calculation, consistent with its standard definition in atmospheric ...
  56. [56]
    [EPUB] Numerical case study of a clear-air turbulence event over the ...
    Feb 24, 2025 · The CAT primarily occurred near the upper-level jet stream core. The significant vertical wind shear decreases the Richardson number (Ri).
  57. [57]
    Clear-Air Turbulence over China: Climatology and Multiscale ... - MDPI
    To improve the positional accuracy, this study uses Automatic Dependent Surveillance–Broadcast (ADS-B) track data (containing time, latitude, longitude, and ...
  58. [58]
  59. [59]
  60. [60]
    [PDF] experimental study of laminar mixed convection in a rod bundle with ...
    However, the Nusselt number varied at the higher elevations as the local Richardson number differed due to an increase in the local heat flux with elevation.
  61. [61]
  62. [62]
  63. [63]
    [PDF] Best Practice Guidelines for the Use of CFD in Nuclear Reactor ...
    Mar 19, 2025 · The emergency core Cooling (ECC) injection systems of the. UPTF were designed to simulate the various ECC systems of PWRs in Germany, Japan.