Fact-checked by Grok 2 weeks ago

Schmidt number

The Schmidt number (Sc) is a dimensionless quantity in fluid mechanics and mass transfer that represents the ratio of a fluid's momentum diffusivity (kinematic viscosity, ν) to its mass diffusivity (D), expressed as Sc = ν / D. This parameter, named after the German engineer Ernst Heinrich Wilhelm Schmidt (1892–1975), quantifies the relative rates at which momentum and mass are transported through a fluid medium, thereby influencing the structure of velocity and concentration boundary layers in convective processes. Physically, a high Schmidt number indicates that mass diffuses much more slowly than momentum, leading to a thinner concentration boundary layer compared to the velocity boundary layer, which is particularly relevant in liquids where molecular diffusion dominates over turbulent mixing. It serves as the mass transfer analog to the Prandtl number in heat transfer, enabling similar scaling analyses for coupled transport phenomena. Typical values of the Schmidt number vary by fluid and conditions: for gases such as air at standard temperature and pressure, Sc is approximately 0.7 to 1, reflecting comparable momentum and mass diffusion rates; in liquids like water, Sc ranges from 100 to 10,000 or higher, due to the much lower mass diffusivity relative to viscosity. In practical applications, the Schmidt number is essential for predicting mass transfer rates in processes such as chemical reactions, pollutant dispersion in environmental flows, gas absorption in oceans, and corrosion in engineering systems, where it helps determine whether diffusion or convection controls solute transport. For turbulent flows, an effective turbulent Schmidt number (often around 0.7–1.0) extends its utility in computational fluid dynamics models to account for enhanced mixing.

Definition and Properties

Mathematical Formulation

The Schmidt number (Sc) is defined as the ratio of the kinematic viscosity (ν) to the molecular mass diffusivity (D) of a species in a fluid, expressed mathematically as
\mathrm{Sc} = \frac{\nu}{D}.
This dimensionless quantity characterizes the relative rates of momentum and mass diffusion in fluid flows.
The kinematic viscosity ν represents the diffusivity of momentum and is given by ν = μ / ρ, where μ is the dynamic viscosity (with units Pa·s) and ρ is the fluid density (with units kg/m³); thus, ν has units of m²/s. The molecular mass diffusivity D quantifies the rate at which mass (e.g., a chemical species) diffuses through the fluid due to concentration gradients and also carries units of m²/s. A units analysis confirms the dimensionless nature of Sc, as both ν and D share the same dimensions (length²/time), making Sc dimensionless. The Schmidt number emerges naturally during the nondimensionalization of the governing equations in fluid dynamics, specifically the incompressible Navier-Stokes equations for momentum conservation and the species conservation (diffusion) equation. To nondimensionalize, characteristic scales are introduced: a reference velocity V, length L, time L/V, concentration c₀, and the equations are scaled accordingly. In the nondimensional Navier-Stokes equation, the viscous diffusion term becomes (1/Re) ∇² u, where Re = V L / ν is the Reynolds number and ∇* is the nondimensional gradient operator. Similarly, in the nondimensional species conservation equation, ∂c*/∂t* + u* · ∇* c* = (1/Pe) ∇² c, where Pe = V L / D is the Péclet number for mass transfer and c* is the nondimensional concentration. The Schmidt number then appears as Sc = ν / D = Pe / Re, linking the relative strengths of the diffusive terms in the two equations.

Physical Interpretation

The Schmidt number represents the ratio of the momentum diffusivity (kinematic viscosity, ν) to the mass diffusivity (D) in a fluid, quantifying how much faster momentum diffuses compared to mass or species in processes involving convective mass transfer. This ratio indicates the relative rates at which velocity perturbations (momentum) spread through the fluid versus how quickly dissolved species or solutes disperse, providing insight into the dominance of viscous effects over molecular diffusion in mass transport phenomena. In boundary layer flows, the Schmidt number governs the relative thicknesses of the velocity (hydrodynamic) boundary layer and the concentration boundary layer. A high Schmidt number implies that mass diffusivity is much lower than momentum diffusivity, resulting in a thinner concentration boundary layer compared to the velocity boundary layer, where species gradients are steeper and confined closer to the surface. Conversely, a low Schmidt number leads to broader concentration boundary layers, as mass diffuses more readily relative to momentum. Typical Schmidt number values vary significantly between gases and liquids due to differences in molecular diffusion rates. For gases like air, Sc is approximately 0.7 under standard conditions, reflecting comparable momentum and mass diffusion. In liquids such as water, Sc is much higher, often around 1000 for common solutes, owing to the substantially lower mass diffusivity in denser fluids. These values highlight that Sc >> 1 for most liquids, emphasizing slower species spreading relative to momentum. In simple laminar or turbulent flows, the Schmidt number qualitatively influences mass transfer coefficients by scaling the Sherwood number in correlations like Sh = f(Re, Sc), where higher Sc typically reduces the effective mass transfer rate for a given flow regime, as the limited molecular diffusion hinders solute transport across boundary layers.

Comparison to Prandtl Number

The Prandtl number, denoted as Pr, is defined as the ratio of momentum diffusivity (kinematic viscosity, ν) to thermal diffusivity (α), expressed mathematically as \Pr = \frac{\nu}{\alpha}. The Schmidt number (Sc) and Prandtl number share a fundamental analogy in transport phenomena, as both quantify the relative rates of momentum diffusion to another diffusive process—mass diffusion for Sc and thermal diffusion for Pr—thereby indicating the relative thicknesses of velocity and concentration/thermal boundary layers in convective flows. This similarity underpins the Reynolds analogy, which assumes equal turbulent transport rates for momentum, heat, and mass when Pr ≈ Sc ≈ 1, allowing simplified predictions of transfer coefficients in high-Reynolds-number flows. Key differences arise in their applications: Sc governs mass transfer processes involving species concentration gradients, whereas Pr pertains to heat transfer driven by temperature gradients. Typical values reflect these distinctions; for gases, Pr and Sc are both approximately 0.7–1.0, leading to comparable boundary layer behaviors, while in liquids, Pr is often around 1–10 (e.g., ~7 for water at room temperature), but Sc is significantly higher at 200–1500 due to slower mass diffusion relative to momentum. This analogy extends to the Chilton-Colburn j-factor method, which correlates friction factors (f/2) with heat and mass transfer Stanton numbers via j_H = \St \Pr^{2/3} and j_m = \St_m \Sc^{2/3}, enabling prediction of mass or heat transfer coefficients from momentum transfer data across a wide range of Pr and Sc values (typically 0.6–10,000).

Relation to Lewis Number

The Lewis number (Le) is a dimensionless parameter that quantifies the relative diffusion rates of heat and mass in fluid systems undergoing simultaneous thermal and solutal transport. It is defined as the ratio of thermal diffusivity \alpha to species mass diffusivity D, \mathrm{Le} = \frac{\alpha}{D}. This relation connects directly to the Schmidt number (Sc) and Prandtl number (Pr) through \mathrm{Le} = \frac{\mathrm{Sc}}{\mathrm{Pr}}, where Sc measures momentum-to-mass diffusivity and Pr measures momentum-to-thermal diffusivity; in certain contexts, such as combustion modeling, the reciprocal form \mathrm{Le} = \mathrm{Pr}/\mathrm{Sc} is employed to emphasize mass diffusion dominance. Physically, the Lewis number represents the comparative scales of thermal diffusion versus molecular diffusion of a species, influencing how temperature and concentration profiles evolve in coupled transport scenarios. For many gases, such as air-water vapor mixtures, Le is approximately 1 due to similar orders of magnitude for \alpha and D (typically $10^{-5} m²/s), implying aligned heat and mass boundary layers. In liquids, however, D is orders of magnitude smaller than \alpha (e.g., D \sim 10^{-9} m²/s versus \alpha \sim 10^{-7} m²/s for water), yielding Le ≫ 1 and indicating faster thermal equilibration relative to solute spreading. In applications involving concurrent heat and mass transfer, such as evaporation from liquid surfaces or premixed combustion, the Lewis number governs the alignment of thermal and concentration fields; when Le ≈ 1, these fields scale similarly, enabling analogies between heat and mass transfer coefficients, whereas Le ≠ 1 introduces differential diffusion effects that can alter flame stability or evaporation rates. For instance, in combustion, Le > 1 stabilizes flames by promoting heat release over fuel diffusion losses, while Le < 1 (common for light fuels like hydrogen in air) can induce cellular instabilities. These insights are critical for designing processes like spray drying or gas turbine combustors where coupled transport dictates efficiency. The Lewis number is named after Warren K. Lewis (1882–1975), the founding head of chemical engineering at MIT, whose seminal 1920s analyses of evaporation and drying processes first highlighted the interplay of heat and mass transfer, influencing fields like psychrometrics (e.g., wet-bulb thermometry assuming Le ≈ 1 for air-water systems) and industrial drying operations.

Turbulent Schmidt Number

Definition and Formulation

The turbulent Schmidt number, denoted Sc_t, quantifies the ratio of turbulent momentum diffusivity to turbulent mass diffusivity in fluid flows dominated by turbulence. It is mathematically formulated as Sc_t = \frac{\nu_t}{D_t} = \frac{\varepsilon}{\varepsilon_M}, where \nu_t represents the eddy kinematic viscosity, responsible for the turbulent transport of momentum, and D_t denotes the eddy diffusivity for mass, governing the turbulent dispersion of scalar species. This definition arises from the need to close the transport equations for both momentum and mass in turbulent flows, assuming analogous gradient-diffusion mechanisms for Reynolds stresses and scalar fluxes. In the framework of Reynolds-averaged Navier-Stokes (RANS) equations, the turbulent Schmidt number emerges during the modeling of turbulent correlations. The RANS momentum equation incorporates the Boussinesq approximation for Reynolds stresses, -\overline{u_i' u_j'} \approx \nu_t \left( \frac{\partial U_i}{\partial x_j} + \frac{\partial U_j}{\partial x_i} \right) - \frac{2}{3} k \delta_{ij}, where U_i is the mean velocity, k is turbulent kinetic energy, and primes denote fluctuations. Similarly, the RANS species transport equation models the turbulent scalar flux as -\overline{u_i' c'} = D_t \frac{\partial C}{\partial x_i}, with C as the mean species concentration. The relation D_t = \nu_t / Sc_t thus provides the necessary closure, linking mass transfer to the established eddy viscosity from turbulence models like k-ε. This formulation preserves the dimensionless character of Sc_t, analogous to the molecular Schmidt number, but shifts its determination from molecular properties (viscosity and diffusivity ratios) to the structural features of the turbulent flow, such as length and time scales of eddies. Unlike the molecular Schmidt number, which varies widely (e.g., approximately 0.7 for gases like air but 600–1000 for liquids like water at room temperature), the turbulent Schmidt number typically assumes values between 0.7 and 1.0 in standard engineering models to reflect near-equivalent turbulent transport efficiencies for momentum and mass. These values are empirically calibrated for simplicity in simulations, though actual Sc_t can deviate based on flow regime, buoyancy effects, or near-wall behavior, highlighting its dependence on turbulence intensity rather than intrinsic fluid properties.

Modeling Approaches

In Reynolds-Averaged Navier-Stokes (RANS) simulations, particularly those employing the standard k-ε turbulence model, the turbulent Schmidt number (Sc_t) is commonly assumed to be a constant value, typically in the range of 0.7 to 0.9, to close the scalar transport equations under the eddy viscosity approximation. This assumption simplifies the modeling of turbulent mass diffusivity by relating it directly to the turbulent viscosity, with many commercial CFD packages defaulting to Sc_t = 0.7 for passive scalar transport in free shear flows. However, this fixed value often underperforms in complex geometries or non-canonical flows, where sensitivity analyses reveal optimal constants varying by up to 50% to match experimental spreading rates. Advanced modeling approaches rely on the gradient diffusion hypothesis, which posits that the turbulent scalar flux is proportional to the mean scalar gradient, analogous to molecular diffusion but scaled by an eddy diffusivity. Under this hypothesis, Sc_t is defined as the ratio of turbulent momentum diffusivity (eddy viscosity) to turbulent mass diffusivity, frequently set equal to the turbulent Prandtl number (Pr_t) for simplicity, as both quantify analogous non-dimensional ratios in momentum and scalar transport. In large eddy simulations (LES), variations of this approach incorporate subgrid-scale models where Sc_t is dynamically adjusted based on local flow resolvability, often yielding values closer to 0.4-0.7 in resolved high-Reynolds-number turbulence to account for anisotropic subgrid fluxes. These methods improve accuracy over constant assumptions by capturing intermittency in scalar mixing, though they increase computational demands. Empirical correlations for Sc_t often express its dependence on the molecular Schmidt number (Sc) or Reynolds number (Re), particularly in high-Sc flows where molecular effects dominate near interfaces. For instance, direct numerical simulations across wide parameter ranges (Re_λ from 8 to 650, Sc from 1/2048 to 1024) indicate that Sc_t exhibits a unique functional form when plotted against the molecular Péclet number (Pe = Re Sc), collapsing scatter observed in individual Re or Sc dependencies and typically yielding Sc_t ≈ 0.6 for isotropic turbulence. In practical engineering contexts, such as jet-in-crossflow configurations, correlations show Sc_t increasing mildly with momentum flux ratios (e.g., from 0.7 to 1.0) or decreasing with Sc in boundary layers, enabling tuned predictions for mass transfer coefficients without full DNS. Modeling Sc_t remains challenging in near-wall regions, where low-Reynolds-number effects and damping of turbulence lead to significant uncertainties, often requiring wall functions or hybrid models to avoid overprediction of scalar gradients. In multiphase flows, such as sediment-laden or particle-dispersed systems, Sc_t deviates from single-phase assumptions due to interphase momentum transfer, with values exceeding 1.0 in non-dilute suspensions from reduced effective diffusivity. Typical literature ranges for Sc_t span 0.4 to 1.5 across environmental and engineering applications, reflecting these sensitivities and underscoring the need for flow-specific calibrations.

Applications

Laminar Flow Mass Transfer

In laminar flow mass transfer, the Schmidt number (Sc) plays a central role in the Sherwood number (Sh), which quantifies the ratio of convective to diffusive mass transfer and is typically expressed as a function of the Reynolds number (Re) and Sc, such as Sh = f(Re, Sc). For mass transfer from a flat plate in a laminar boundary layer, the average Sherwood number over the plate length L is given by
Sh_{L,avg} = 0.664 \, Re_L^{1/2} \, Sc^{1/3},
where Re_L is based on the plate length and free-stream velocity, valid for dilute solutions and Sc > 0.6. This correlation arises from solving the convective diffusion equation using similarity transformations analogous to the Blasius solution for momentum boundary layers, highlighting Sc's influence on the concentration boundary layer thickness.
Higher values of Sc result in steeper concentration gradients near the surface because the molecular diffusivity is low relative to momentum diffusivity, leading to thinner concentration boundary layers and consequently lower mass transfer coefficients for a given flow rate. In external flows like the flat plate, this manifests as a weaker dependence of Sh on Re compared to low-Sc cases, where diffusion dominates more broadly. For internal flows, such as in ducts, the effect is pronounced in developing regions, where high Sc confines mass transfer to a narrow layer adjacent to the wall, reducing overall rates unless enhanced by flow development. Analytical solutions for laminar mass transfer in developing flows, such as the Graetz problem for circular ducts, provide exact expressions for the Sherwood number as a function of the Graetz number (Gz = Re Sc d/L, where d is the diameter and L the axial length). The Graetz series solution yields Sh approaching 3.66 for fully developed concentration profiles under constant wall concentration, but in the entrance region (high Gz), it scales with Sc through the diffusion equation eigenvalues. For high Sc, the Lévêque approximation simplifies this by assuming a linear velocity profile near the wall, leading to the local Sherwood number scaling as Sh_x ∝ (Re Sc / (x/d))^{1/3}, where x is the axial distance. Experimental validations of these correlations often employ electrochemical techniques to measure mass transfer in pipes and channels at high Sc (e.g., Sc ≈ 1000 for aqueous electrolytes), confirming the Lévêque form for entry lengths where Gz > 100. A widely adopted average Sherwood number correlation for laminar developing flow in circular tubes is
Sh_{avg} = 1.62 \, (Re \, Sc \, d/L)^{1/3},
accurate to within 5-10% for short tubes (L/d < 0.1) and high Sc, as verified against dissolution and limiting current density data. In parallel-plate channels, similar forms hold with adjusted constants, emphasizing Sc's role in amplifying entrance effects for low-diffusivity solutes.

Turbulent Flow Simulations

In computational fluid dynamics (CFD) simulations of turbulent flows, the turbulent Schmidt number (Sc_t) is integrated into Reynolds-Averaged Navier-Stokes (RANS) turbulence models to predict scalar dispersion and mass transfer. For instance, in the k-ω Shear Stress Transport (SST) model, Sc_t modulates the eddy diffusivity for momentum and scalars, enabling accurate forecasting of concentration fields in flows with adverse pressure gradients or separation. This incorporation is crucial for applications involving passive scalar transport, where a typical Sc_t value around 0.85 is often tuned based on flow specifics to balance numerical stability and physical fidelity. For high Schmidt number (Sc > 100) regimes, common in liquid-phase mass transfer, simulations require enhanced near-wall resolution to capture the thin diffusive sublayer, significantly affecting the accuracy of both Direct Numerical Simulation (DNS) and RANS approaches. In DNS, high Sc leads to steeper concentration gradients near walls, necessitating finer grids to resolve molecular diffusion without excessive computational cost, as demonstrated in channel flow studies where insufficient resolution underpredicts mass transfer rates by up to 20%. RANS models, reliant on Sc_t for subgrid-scale closure, exhibit sensitivity to near-wall treatments, with low-Re variants like low-Reynolds k-ω improving predictions but still demanding y+ < 1 for Sc > 100 to avoid overestimation of Sherwood numbers. Case studies illustrate Sc's role in mixing efficiency across diverse turbulent scenarios. In atmospheric pollutant dispersion, low Sc (around 0.7-1 for gases) promotes rapid turbulent mixing, as seen in urban street canyon simulations where optimized Sc_t values enhance plume spread predictions compared to fixed defaults, influencing exposure assessments. Conversely, in chemical reactors with high Sc (e.g., >1000 for electrolytes), Sc governs the dominance of molecular diffusion in boundary layers, reducing overall mixing efficiency and requiring Sc_t adjustments in RANS to match experimental dilution rates, highlighting slower scalar homogenization in liquid systems. Validation of these simulations often involves comparing computed Sherwood numbers (Sh) against experimental data, underscoring Sc_t's sensitivity. For example, in impinging jet studies, RANS predictions of Sh align within 10% of measurements when Sc_t is varied from 0.7 to 1.0, but diverge significantly for high Sc due to inadequate turbulence-scalar coupling. Such metrics reveal that Sc_t overprediction can inflate Sh by 15-25% in wall-bounded flows, emphasizing the need for case-specific calibration to ensure reliable turbulent mass transfer forecasts.

Stirling Engine Analysis

In Stirling engines, the working fluid typically consists of gases like helium, which exhibits a Schmidt number of approximately 0.7 at standard conditions, reflecting the comparable rates of momentum and mass diffusivity in low-density gases. This value is characteristic of gaseous working fluids used in the regenerator and heat exchangers, where oscillatory flows dominate the transport processes. The low Schmidt number implies relatively thick diffusive boundary layers, influencing the efficiency of heat storage and recovery in these components. The Schmidt number significantly impacts axial dispersion within the regenerator, where the periodic motion of gas parcels can lead to enhanced mixing due to molecular diffusion relative to viscous effects. In oscillatory flows, this dispersion reduces the sharpness of temperature fronts, contributing to imperfect regeneration and lower cycle efficiency, particularly in compact designs with high surface-area-to-volume ratios. For helium-based systems, the Schmidt number's role becomes prominent at low Reynolds numbers typical of Stirling operation, where diffusive spreading can account for up to several percent of performance losses. Representative studies show that optimizing regenerator porosity and length can mitigate these effects, improving overall engine effectiveness by 5-10% in simulated conditions. In one-dimensional modeling of Stirling engines, the Schmidt number is integrated into calculations for pressure drop and shuttle losses, often through correlations linking it to the Sherwood number for mass transfer analogies in dead volumes and appendage spaces. These models account for charging and discharging inefficiencies during the cycle, where Sc influences the Nusselt-Sherwood analogy for combined heat and momentum transfer in oscillatory regimes; for instance, Sherwood numbers derived from Sc-based correlations help quantify frictional dissipation in the compression and expansion spaces. Such approaches enable predictions of net power output, with Sc-dependent terms adjusting for viscous and diffusive contributions to entropy generation. Gustav Schmidt's seminal 1871 analysis of the ideal Stirling cycle, assuming isothermal compression and expansion with sinusoidal piston motion, implicitly incorporated volume and capacity ratios that foreshadow modern dimensionless groupings like the Schmidt number in loss-inclusive models. This foundational work established the baseline efficiency expression, later refined with Sc to address real-fluid dispersive effects in regenerator performance.

Other Engineering Contexts

In environmental engineering, the Schmidt number plays a crucial role in modeling the diffusion of carbon dioxide across ocean surfaces, where it quantifies the ratio of momentum diffusivity to molecular diffusivity for CO₂ in seawater, typically yielding a value of approximately 660 at 20°C and standard salinity. This parameter is essential for parameterizing air-sea gas transfer velocities in global carbon cycle models, as it influences the scaling of transfer rates with wind speed and temperature variations in oceanic boundary layers. For atmospheric trace gases, such as methane or volatile organic compounds, the turbulent Schmidt number in air flows is generally around 0.7 to 1.0, reflecting the near-unity ratio of turbulent viscosity to eddy diffusivity near the surface layer, which affects emission flux estimates from environmental sources like soils or water bodies. In chemical engineering, high Schmidt numbers are prevalent in membrane separation processes, where low molecular diffusivities of solutes relative to fluid viscosity (often Sc > 1000 for aqueous systems) lead to thin concentration boundary layers that enhance separation efficiency but require careful hydrodynamic design to mitigate polarization effects. For instance, in ultrafiltration or reverse osmosis, the Schmidt number informs Sherwood number correlations to predict mass transfer coefficients, directly impacting permeate flux and rejection rates for macromolecules. In electrochemical processes, such as electrodeposition or electrolysis, elevated Schmidt numbers (typically 500–50,000 in viscous electrolytes) control the diffusion layer thickness, thereby influencing reaction selectivity by limiting reactant transport to the electrode surface and favoring specific ion pathways over competing reactions. Biomedical applications leverage the Schmidt number to analyze solute transport in blood plasma, where values range from approximately 10³ to 10⁴ for oxygen and other small solutes due to their low diffusion coefficients (around 10⁻⁵ cm²/s) in the viscous plasma medium, emphasizing diffusive limitations in microcirculatory oxygen delivery. This high Sc regime governs the mass transfer from plasma to tissues or across capillary walls, where it scales the boundary layer resistance in models of oxygen unloading from hemoglobin, critical for understanding hypoxemia or tissue oxygenation in pathological conditions. In aerospace engineering, low Schmidt numbers (around 0.6–0.7) characterize the behavior of ablation products and gases in hypersonic flows, where high thermal diffusivities relative to mass diffusivities facilitate rapid mixing in the boundary layer during re-entry vehicle ablation modeling. This parameter is incorporated into computational fluid dynamics simulations to predict species diffusion and thermochemical nonequilibrium effects, ensuring accurate forecasting of material erosion rates and heat shield performance under extreme conditions.

History and Development

Origin of the Concept

The roots of the Schmidt number concept trace back to the mid-19th century, when Adolf Fick formulated his laws of diffusion in 1855, establishing a mathematical framework for mass transfer analogous to Fourier's law of heat conduction. This foundational work described diffusive flux as proportional to the concentration gradient, providing the essential physical basis for later dimensionless analyses in mass transport processes. Fick's contributions enabled the recognition of molecular diffusivity as a key parameter, setting the stage for integrating mass transfer into broader fluid dynamics studies. Building on this, early 20th-century developments in boundary layer theory and transfer analogies further shaped the concept. Osborne Reynolds introduced ideas on the interaction between momentum and energy transport in his 1874 investigations of fluid motion, proposing an analogy between skin friction and heat transfer in turbulent flows that highlighted the role of viscous effects near surfaces. This Reynolds analogy laid preliminary groundwork for non-dimensionalizing transport phenomena, influencing subsequent work on boundary layers. Paralleling these advances, Ludwig Prandtl formalized the Prandtl number in 1910 through his analysis of fluids with low viscosity, defining it as the ratio of momentum diffusivity to thermal diffusivity and establishing a model for analogous groups in heat transfer. The explicit emergence of the Schmidt number occurred in mass transfer literature during the 1920s, as researchers began applying dimensionless formulations to evaporation and absorption processes, mirroring the Prandtl number's role in heat transfer. Early texts and correlations, such as those drawing from Nusselt's 1915 dimensionless groups for convective transfer, incorporated ratios involving mass diffusivity to generalize empirical data across scales. The term itself was coined in the late 1920s by German engineer Ernst Heinrich Wilhelm Schmidt (1892–1975), who explicitly highlighted the analogy between heat and mass transfer in boundary layers, defining the Schmidt number as the ratio of momentum diffusivity to mass diffusivity. This naming reflected Schmidt's influential 1929 paper "Verdunstung und Wärmeübergang," which solidified the parameter's place in fluid mechanics. Over the course of the 20th century, the Schmidt number evolved from ad hoc empirical correlations in experimental mass transfer studies to a core dimensionless group in theoretical fluid mechanics, facilitating similarity analyses and scaling laws for diverse systems. This progression was driven by advances in understanding turbulent and laminar transport, where the number's value—typically around 1 for gases and higher for liquids—quantifies the relative dominance of viscous versus diffusive effects.

Key Contributions and Evolution

The Schmidt number was first formalized through the work of German engineer Ernst Schmidt, who in his 1929 paper "Verdunstung und Wärmeübergang" emphasized the fundamental analogies between momentum, heat, and mass transfer processes, introducing the dimensionless group that bears his name to characterize mass diffusivity relative to viscous diffusion. Following World War II, the Schmidt number gained prominence in the unified framework of transport phenomena, notably through its detailed exposition in the seminal 1960 textbook Transport Phenomena by R. Byron Bird, Warren E. Stewart, and Edwin N. Lightfoot, which integrated it into broader analyses of multicomponent systems and diffusion. In the 1970s, extensions to turbulent flows advanced the concept, with Kemal Hanjalić and Brian E. Launder incorporating a turbulent Schmidt number into their Reynolds stress model for predicting scalar transport in shear flows, assuming a value around 0.9 for eddy diffusivity modeling. Modern developments in the 1990s leveraged direct numerical simulations (DNS) to validate and reveal the variability of the turbulent Schmidt number across different flow regimes and molecular Schmidt values, as demonstrated in studies of passive scalar mixing in isotropic turbulence, showing dependencies on Reynolds number and scalar gradients. These refinements addressed early limitations, such as the neglect of variable fluid properties in analogies, which are now routinely accounted for in high-accuracy multiphysics simulations to capture effects like density variations in compressible or reacting flows.

References

  1. [1]
    [PDF] A Glossary of Terms for Fluid Mechanics - University of Notre Dame
    Schmidt Number (Sc). Definition: ν. D. D is the mass diffusivity. This number is the ratio of momentum diffusion (kinematic viscosity) to mass diffusion. It ...
  2. [2]
    [PDF] File Fundamentals Of Momentum Heat And Mass Transfer Solutions
    It was named after German engineer Ernst Heinrich Wilhelm Schmidt (1892–1975). The Schmidt number is the ratio of the shear component for diffusivity (viscosity.
  3. [3]
    [PDF] Convective Mass Transfer - Clarkson University
    The Schmidt number plays a role in mass transfer that is analogous to that played by the Prandtl number in heat transfer. From its definition, we can infer a ...
  4. [4]
    [PDF] 3.21 Lectures on Fluid Flow and Kinetics
    The mass transfer Prandtl number ν/D is also called the Schmidt number. ... Biot number, which is defined using the thermal conductivity or diffusivity of a solid ...
  5. [5]
    On the Values for the Turbulent Schmidt Number in Environmental ...
    The Schmidt number Sc is usually in the order of one and of 102–102 depending on temperature for environmental flows in air and water, respectively (Table 1).
  6. [6]
    Effects of Schmidt Number on Turbulent Mass Transfer Around a ...
    In corrosion applications where the fluid is usually a liquid, the Schmidt number (Sc) plays an important role in the determination of the mass transfer rate ...
  7. [7]
    [PDF] Dimensional Analysis in Mass Transfer
    Jan 5, 2021 · z-component of the Navier-Stokes Equation: │. ⎠. ⎞. │. ⎝. ⎛. ∂. ∂. + ... Schmidt number. (a material property). © Faith A. Morrison ...
  8. [8]
    Schmidt Number | Neutrium
    Oct 14, 2018 · The Schmidt number is a dimensionless number named after Ernst Heinrich Wilhelm Schmidt and describes the ratio of momentum diffusivity to ...Missing: history origin
  9. [9]
    Schmidt Number - an overview | ScienceDirect Topics
    The Schmidt number is a dimensionless quantity that relates the viscosity of a fluid to its diffusion coefficient, correlating momentum transport with mass ...
  10. [10]
    Schmidt number | tec-science
    May 10, 2020 · The Schmidt number is a dimensionless similarity parameter to describe mass and momentum transfer. Only with identical Schmidt numbers one obtains physically ...
  11. [11]
    Concentration Boundary Layers | Heat and Mass Transfer Class Notes
    Higher Schmidt numbers indicate a thinner concentration boundary layer relative to the velocity boundary layer ... Schmidt number ($Sc$): Relates the ...
  12. [12]
    Effect of Schmidt number on mass transfer across a sheared gas ...
    Nov 14, 2016 · The mass transfer across a sheared gas-liquid interface strongly depends on the Schmidt number. Here we investigate the relationship between mass transfer ...
  13. [13]
    Prandtl Number - an overview | ScienceDirect Topics
    Prandtl values for some selected fluids. Usually, the Prandtl number is assumed to be around 0.7 for gases and around 6.9 for water.
  14. [14]
    Chilton and Colburn J factor analogy - Chemepedia
    Apr 27, 2020 · Chilton and Colburn J factor analogy · Sc is the Schmidt Number · Pr is the Prandtl Number · Re is the Reynolds Number · Sh is the Sherwood Number.
  15. [15]
    Lewis Number - an overview | ScienceDirect Topics
    The premixed flame is stable when the Lewis number is greater than unity (Le > 1) and unstable for Lewis number less than unity (Le < 1). Moreover, the Lewis ...
  16. [16]
    CALCULATING the LEWIS NUMBER
    Sc denotes the Schmidt number and Pr denotes the Prandtl number. It should be noted that sometimes the Lewis number is defined as the inverse of the ...
  17. [17]
    Calculated Prandtl number, Schmidt number, and Lewis number of ...
    The Schmidt number is the ratio of momentum ͑ viscous ͒ to mass diffusion effects. It remains fairly constant over the range of temperature considered here. It ...
  18. [18]
    Convection patterns in colloidal solutions | Phys. Rev. E
    Mar 6, 2007 · In liquid molecular mixtures a typical value for the Lewis number is 0.01, whereas in colloids L can easily reach values as low as 10 − 4 .
  19. [19]
    Effects of Lewis number on the statistics of the invariants of the ...
    ... Lewis number Le, which is defined as the ratio of thermal diffusivity to mass diffusivity. Although every species in a combustion process has its own Lewis ...
  20. [20]
    Effects of Lewis number on turbulent kinetic energy transport in ...
    Jul 27, 2011 · ... Lewis number flames in which the rates of heat and mass diffusion are equal to each other. The Lewis number is defined as the ratio of ...
  21. [21]
    On the effective Lewis number formulations for lean hydrogen ...
    May 10, 2013 · The Lewis number Le, named in honor of Warren K. Lewis for his pioneering studies on liquid evaporation [1], is nowadays commonly defined as ...
  22. [22]
    On the Centennial of the Academic Careers (1919–2019) of ...
    The subject area of heat and mass transfer began to be formalized for college curricula and classroom instruction, driven by industrial needs and the ...
  23. [23]
    Turbulent Schmidt numbers for CFD analysis with various types of ...
    To represent the ratio of the turbulent momentum diffusivity (eddy viscosity) νt and the turbulent mass diffusivity Dt, the turbulent Schmidt number, Sct is ...Missing: exact formula
  24. [24]
    (PDF) On the Values for the Turbulent Schmidt Number in ...
    Apr 17, 2017 · The turbulent Schmidt number, defined as the ratio of turbulent eddy viscosity to turbulent mass diffusivity, is typically assigned a value ...
  25. [25]
    [PDF] Turbulence Models Applied Computational Fluid Dynamics - Icmc-Usp
    Experiments have shown that the turbulent Schmidt number is nearly constant with typical values between 0.7 and 1. i t i x u. ∂. Φ∂.
  26. [26]
    [PDF] Lecture 10 Turbulent Combustion: The State of the Art
    In deriving the basic LES equations, the Navier-Stokes equations are spatially filtered with a filter of size Δ, which is of the size of the grid cell (or a ...
  27. [27]
  28. [28]
    Turbulent Schmidt numbers for CFD simulations using the k-ε and k ...
    Aug 10, 2025 · Turbulent diffusion dominates mixing at high molecular Schmidt numbers. When solving the Reynolds averaged Navier-Stokes equations, one of the ...<|control11|><|separator|>
  29. [29]
  30. [30]
  31. [31]
    None
    ### Summary of Laminar Boundary Layer Mass Transfer Correlations for Sherwood Number
  32. [32]
    [PDF] THE GENERALIZED LÉVÊQUE EQUATION AND ITS USE TO ...
    The Generalized Lévêque Equation (GLE) is a generalization of the Lévêque equation, used to predict heat or mass transfer from fluid friction, with the form Nu ...
  33. [33]
    Refinement of the Generalized Graetz Problem Correlation With ...
    Mar 17, 2020 · The refined Graetz problem correlation for fully developed flow is able to predict exact thermal entry region solutions to within ±1.2% for all ...
  34. [34]
    [PDF] THE LÉVÊQUE-ANALOGY or HOW TO PREDICT HEAT AND MASS ...
    The Lévêque analogy uses the Generalized Lévêque Equation (GLE) to calculate heat/mass transfer coefficients from frictional pressure drop, not flow rates.
  35. [35]
    [PDF] Theoretical Considerations of Pressure Drop and Mass Transfer of ...
    May 31, 2016 · Re.Sc.dh l. < 104 ; the Sherwood number can be calculated by the very well-known Levêque equation. [22]:. Sh = 1.62. Re.Sc. dh l ! "#. $. %&. 1/ ...
  36. [36]
    [PDF] Theoretical Considerations of Pressure Drop and Mass Transfer of ...
    Mar 2, 2022 · Re.Sc.dh l. < 104 ; the Sherwood number can be calculated by the very well-known Levêque equation. [22]:. Sh = 1.62. Re.Sc. dh l ! "#. $. %&. 1/ ...
  37. [37]
    transport phenomena in membrane separation processes - j-stage
    Sh=1.62(Re Sc dn/L)¹/3. (20). Deissler equation (turbulent flow):. Sh=0.023 Re0 ... Sherwood number temperature mobility of ion convective coupling ...
  38. [38]
    [PDF] On the turbulence models and turbulent Schmidt number in ...
    Stratified flows are prevalent in indoor and outdoor environments. To predict these flows, this investigation evaluated the performance of seven turbulence ...
  39. [39]
    Optimal Turbulent Schmidt Number for RANS Modeling of Trailing ...
    Most RANS simulations use a fixed turbulent Schmidt number of either 0.7 or 0.85 to determine the turbulent scalar flux, based on the values for canonical flows ...
  40. [40]
    Modelling high Schmidt number turbulent mass transfer
    The DNS, which considers passive scalars with Schmidt numbers between 1 and 50, is used to analyse the mass transfer coefficient K and the near-wall behaviour ...
  41. [41]
    High-Schmidt-number mass transport mechanisms from a turbulent ...
    Aug 6, 2012 · We have investigated the mechanisms involved in dissolved oxygen (DO) transfer from a turbulent flow to an underlying organic sediment bed ...
  42. [42]
    Sensitivity of the turbulent Schmidt number and the ... - ResearchGate
    This work presents numerical simulations, with validation considering analytical expressions and experimental results, of masstransfer in electrochemical ...
  43. [43]
    Mass transfer in turbulent impinging slot jets - ScienceDirect.com
    The mass transfer characteristics of a turbulent slot jet impinging normally on a target wall are examined using numerical simulations.
  44. [44]
  45. [45]
    [PDF] Stirling Engine Design Manual
    Mar 27, 1978 · This is a Stirling Engine Design Manual, second edition, prepared for NASA, covering what a Stirling engine is and major types.
  46. [46]
    [PDF] Schmidt analysis for Stirling Engines
    The difference between the two analysis is that the Rejector (k) in the engine is the lower temperature reject sink and in the cooler is the higher temperature ...
  47. [47]
    Relationship between wind speed and gas exchange over the ...
    Jun 5, 2014 · Empirical relationships of the Schmidt number, which are necessary to determine the fluxes, are extended to 40°C to facilitate their use in the ...
  48. [48]
    Turbulent Schmidt number from a tracer experiment - ScienceDirect
    The value of Sc has implications for the measurement of trace gas emissions, and there is a broad range of reported values for the atmospheric surface layer.
  49. [49]
    high Schmidt mass transfer predictions in laminar flow - ScienceDirect
    A correlation is presented between stagnation Sherwood number, jet Reynolds number and Schmidt number, Shst/Sc1/3=1.771/2. The effects on the mass transfer ...
  50. [50]
    Electrochemical Mass Transfer Measurements with Glycerin Used ...
    After optimization of the experimental conditions, a Schmidt number of about 50 000 was obtained, higher than those attained previously and likely the highest ...
  51. [51]
    Ionic mass transfer in parallel plate electrochemical cells
    For the turbulent region, there is a correlation of the mass transfer coefficient with Reynolds number to an exponent of 0.875 and Schmidt number to exponent ...Missing: selectivity | Show results with:selectivity
  52. [52]
    Pulsatile blood flow and oxygen transport past a circular cylinder
    ... Schmidt number (Sc). The Hill equation is used to describe the saturation ... For a pure O(2) feed, oxygen transport in the plasma dominates near the cylinder.
  53. [53]
    [PDF] CFD Analysis of Hypersonic Flowfields with Surface ...
    Schmidt number based mass diffusion coefficients and temperature depen- dent ... solutionswith coupled ablation. This has been done in the design phase ...
  54. [54]
    [PDF] Hypersonic Flow Analysis Including Finite Rate Ablation ... - eucass
    The diffusion model is based on an effective diffusion coefficient obtained assuming a constant Schmidt number of 0.7. The gas-phase chemical reactions ...
  55. [55]
    Fick's Law - an overview | ScienceDirect Topics
    Fick's law refers to two fundamental laws of diffusion that describe how diffusive flux is related to concentration gradients. The first law states that the ...Missing: dimensionless | Show results with:dimensionless
  56. [56]
    HISTORY OF BOUNDARY LA YER THEORY - Annual Reviews
    HISTORY OF BOUNDARY-LAYER THEORY. 103 was noticed by Prandtl (1910) and Taylor (1916). Effects of waH roughness were also discussed with consideration for ...
  57. [57]
    ORIGINS OF DIMENSIONLESS GROUPS OF HEAT AND MASS ...
    A paper pointing out the analogy between heat and mass transfer caused the dimensionless quantity involved to be called "Schmidt Number". In 1937 he became ...
  58. [58]
    Ernst Schmidt (1892-1975)
    A paper pointing out the analogy between heat and mass transfer caused the dimensionless quantity involved to be called the "Schmidt Number." In 1937, he ...
  59. [59]
    [PDF] Notes on the Origins and Evolution of the Subject of Heat Transfer
    German contribu- tions in convection, conduction radiation, and heat ex- changer design continued through the 1920s almost as though World War I and the ...