Fact-checked by Grok 2 weeks ago

Lewis number

The Lewis number (Le) is a dimensionless parameter in and that quantifies the relative rates of thermal diffusion and mass diffusion in a , defined as the ratio of (α) to (D), expressed as Le = α / D. It can also be formulated as Le = Sc / Pr, where Sc is the and Pr is the , providing a measure of how and boundary layers develop in convective flows. Named after Warren K. Lewis (1882–1975), a pioneering often regarded as the father of in the United States, the number was introduced to characterize in systems involving simultaneous and . In engineering applications, the Lewis number plays a critical role in analyzing and predicting behavior in processes such as , where it assesses the stability of premixed flames: values greater than unity ( > 1) typically indicate stable flames due to faster relative to , while < 1 can lead to thermal-diffusive instabilities, as seen in lean hydrogen mixtures ( ≈ 0.437). For air and many common gases like methane ( ≈ 0.999), the value is near 1, implying comparable and concentration boundary layer thicknesses, which simplifies modeling in exchangers and dehumidification systems. In turbulent flows, such as those in chemical reactors or HVAC designs, deviations from = 1 influence coupled and transfer efficiency, guiding material selection and geometry optimization for enhanced performance. The parameter is particularly valuable for scaling experiments, as systems with identical Lewis numbers exhibit analogous transport behaviors regardless of size.

Definition

Physical Interpretation

The Lewis number (Le) quantifies the relative rates at which heat and mass diffuse within a fluid, serving as the ratio of thermal diffusivity (α), which describes how quickly heat spreads through conduction, to mass diffusivity (D), which describes the analogous spread of chemical species via molecular diffusion. This parameter captures the comparative efficacy of these two transport mechanisms in scenarios involving coupled heat and mass transfer, such as in boundary layer flows where temperature and concentration gradients coexist. Physically, the value of Le reveals the dominance of one diffusion process over the other, influencing the structure of associated s without incorporating momentum transport effects like viscous diffusion. When Le > 1, diffusion prevails over mass diffusion, resulting in a thicker and a correspondingly thinner concentration . In the opposite case, Le < 1 signifies faster mass diffusion, yielding a thinner relative to the concentration one; Le ≈ 1 denotes balanced diffusion rates, with similar thicknesses for both s. Representative values underscore these trends across fluid states: Le ≈ 1 is common for many gases, including air-fuel mixtures like stoichiometric methane-air, where and diffuse at comparable speeds. Liquids, however, typically exhibit Le >> 1—around 100 for —due to the markedly slower diffusion in dense media compared to thermal . The Lewis number thus provides a momentum-independent lens for analyzing relative and transport, distinct from analogs like the Prandtl and numbers that incorporate viscous effects.

Mathematical Formulation

The Lewis number Le, a in heat and mass transfer, is defined as the ratio of the \alpha to the D of a mixture: Le = \frac{\alpha}{D} where the is given by \alpha = \frac{k}{\rho c_p}, with k denoting the thermal conductivity, \rho the mixture density, and c_p the at constant pressure. This formulation arises from the governing equations for energy and species conservation, capturing the relative rates of and mass . Substituting the expression for yields the expanded form: Le = \frac{k}{\rho c_p D} Since both \alpha and D share units of length squared per unit time (m²/s), the Lewis number is inherently dimensionless, facilitating scale-independent analysis in . An equivalent expression relates the Lewis number to other fundamental dimensionless groups in : the Sc = \frac{\nu}{D} (ratio of momentum diffusivity to , with \nu as kinematic viscosity) and the Pr = \frac{\nu}{\alpha} (ratio of momentum diffusivity to thermal diffusivity): Le = \frac{Sc}{Pr} This relation highlights the Lewis number's role as a bridge between momentum, heat, and mass transfer characteristics. In binary mixtures, D represents the binary mass diffusivity between the two species, providing a straightforward measure; for multi-component systems, D is typically approximated using a binary or mixture-averaged diffusivity to simplify calculations while maintaining accuracy for dominant species interactions.

Comparison with Other Dimensionless Numbers

The Lewis number (Le) relates to other dimensionless numbers in by expressing the ratio of to , specifically Le = Sc / Pr, where Sc is the and Pr is the . This relation positions Le as a measure of relative diffusive transport excluding momentum effects, in contrast to Pr and Sc, which both incorporate kinematic viscosity (ν). The is defined as = ν / α, comparing diffusivity (ν) to (α); for high (typical in oils or viscous fluids, > 1), diffuses faster than , resulting in thinner thermal boundary layers embedded within thicker boundary layers. Similarly, the is = ν / D, comparing diffusivity to (D); high (common in liquids, >> 1) implies slower mass diffusion, leading to thin concentration boundary layers relative to the momentum layer. Unlike and , which highlight viscous influences on transport, isolates the balance between thermal and mass diffusion, making it particularly useful in scenarios where or reaction rates depend on these diffusive ratios without direct coupling. When ≈ 1, as in many air-fuel mixtures, and diffusivities are comparable, rendering and processes analogous and allowing simplified modeling by equating developments or flux expressions. This unity Lewis number approximation streamlines numerical simulations and analytical solutions in or problems by reducing the need for separate treatment of and species equations. However, deviations from Le = 1 (e.g., Le < 1 in hydrogen s or Le > 1 in heavy hydrocarbons) introduce differential diffusion effects, which can alter structures or transfer rates and thus impact model accuracy if neglected. Le also influences related convective transport numbers, such as the Peclet numbers for (Pe_h = Re Pr = uL / α) and (Pe_m = Re Sc = uL / D), where the ratio Pe_h / Pe_m = D / α = 1 / Le quantifies the relative dominance of over conduction versus in and contexts.

Variations and Extensions

In contexts, the Lewis-Semenov number (Le_S), named after Warren K. and Semenov, is defined as Le_S = \frac{\lambda}{\rho c_p D}, where \lambda is the thermal conductivity, \rho is the , c_p is the at constant pressure, and D is the binary mass diffusivity. This formulation arises in analyses of reactive flows where it quantifies the balance between heat conduction and , often assuming unity or specific scaling in non-dimensional equations, and is sometimes used interchangeably with the standard Lewis number despite subtle distinctions in derivation for flame propagation and . The inverse Lewis number (1/Le) appears in literature focused on mass-transfer-dominated regimes, such as drying processes, where it emphasizes scenarios in which mass diffusion outpaces thermal diffusion, altering boundary layer development and transfer rates. For instance, in viscoelastic liquid films or solute transport models, 1/Le = D / \alpha serves as a key parameter to assess instability and phase separation dynamics. For multi-component mixtures, particularly in of fuel blends, an effective Lewis number (Le_eff) is employed to approximate the overall diffusional-thermal behavior. A widely adopted is 1 / Le_eff = \sum_i Y_i / Le_i, where Y_i denotes the of component i and Le_i its individual Lewis number; this weighted accounts for varying diffusivities across , influencing and extinction limits in surrogates like hydrocarbon-air mixtures. The Lewis number exhibits dependence on and , reflecting changes in properties. In gases, Le is approximately constant with rising temperature, as both the D and the thermal diffusivity \alpha scale similarly with T (around T^{1.7} to T^{1.8} depending on the gas), though slight variations occur; pressure effects inversely scale both diffusivities, often maintaining approximate constancy at fixed T but amplifying variations in non-ideal conditions. Definitions of the Lewis number show inconsistencies across disciplines. In some fields like , particularly in analyses of ion and effects in electrolytes, Le is defined inversely as D / \alpha to prioritize over in coupled electrochemical- models.

Derivation

From Governing Equations

The derivation of the Lewis number begins with the fundamental governing equations for heat and mass transfer in a fluid flow, assuming incompressible flow with constant properties and neglecting source terms for simplicity. The energy equation for temperature T is given by \frac{\partial T}{\partial t} + \mathbf{u} \cdot \nabla T = \alpha \nabla^2 T, where \mathbf{u} is the velocity vector and \alpha = k / (\rho c_p) is the thermal diffusivity, with k the thermal conductivity, \rho the density, and c_p the specific heat capacity. The corresponding species transport equation for concentration C (e.g., of a scalar species) is \frac{\partial C}{\partial t} + \mathbf{u} \cdot \nabla C = D \nabla^2 C, where D is the mass diffusivity. To reveal dimensionless parameters, non-dimensionalize these equations using characteristic scales: length L, velocity U (implying time scale L/U), temperature difference \Delta T, and concentration difference \Delta C. Define dimensionless variables as \hat{\mathbf{x}} = \mathbf{x}/L, \hat{t} = t U / L, \hat{\mathbf{u}} = \mathbf{u}/U, \hat{T} = (T - T_0)/\Delta T, and \hat{C} = (C - C_0)/\Delta C, where T_0 and C_0 are reference values. Substituting into the energy equation yields \frac{\partial \hat{T}}{\partial \hat{t}} + \hat{\mathbf{u}} \cdot \nabla \hat{T} = \frac{1}{\mathrm{Pe}_h} \nabla^2 \hat{T}, where the hats are dropped for brevity and \mathrm{Pe}_h = U L / \alpha is the thermal Péclet number, representing the ratio of convective to diffusive heat transport. Similarly, the species equation becomes \frac{\partial \hat{C}}{\partial \hat{t}} + \hat{\mathbf{u}} \cdot \nabla \hat{C} = \frac{1}{\mathrm{Pe}_m} \nabla^2 \hat{C}, with \mathrm{Pe}_m = U L / D the mass Péclet number. The convective terms scale identically in both equations (order U \Delta T / L or U \Delta C / L), so they cancel upon normalization, leaving the diffusive terms balanced by the inverse Péclet numbers. In coupled heat and mass transfer problems, the relative magnitudes of the diffusive contributions determine the interaction between thermal and species fields. The ratio of the thermal diffusive term \alpha \nabla^2 T to the mass diffusive term D \nabla^2 C (scaled appropriately) introduces the Lewis number as \mathrm{Le} = \alpha / D = \mathrm{Pe}_m / \mathrm{Pe}_h, which quantifies the balance between thermal diffusion and mass diffusion. This parameter governs how temperature and concentration profiles evolve together under convection, emerging as the key coefficient linking the two normalized equations. For instance, rewriting the species equation in terms of the energy scaling gives a diffusive coefficient of $1/(\mathrm{Pe}_h \mathrm{Le}), highlighting Le's role in diffusive disparity. In boundary layer flows, such as those analyzed via similarity solutions (e.g., the Blasius solution for the momentum boundary layer), the Lewis number further manifests in the relative thicknesses of the thermal and concentration boundary layers. Order-of-magnitude scaling shows the thermal boundary layer thickness \delta_\mathrm{thermal} \sim \sqrt{\alpha x / U} and the concentration boundary layer thickness \delta_\mathrm{concentration} \sim \sqrt{D x / U}, where x is the streamwise distance. Thus, the ratio \delta_\mathrm{thermal} / \delta_\mathrm{concentration} \approx \sqrt{\mathrm{Le}}, indicating that for Le > 1 (faster thermal diffusion), the thermal layer is thicker than the concentration layer. This scaling arises directly from the diffusive balance in the non-dimensional equations, with convective terms again canceling to leave Le as the governing parameter for profile similarity.

Assumptions and Limitations

The derivation of the Lewis number from the governing equations for and assumes constant thermophysical properties, specifically that the \alpha and D remain independent of T and concentration C. This simplification facilitates the ratio \mathrm{Le} = \alpha / D but holds primarily under isothermal or mildly varying conditions. Additionally, the formulation presumes low flows, treating the fluid as incompressible to neglect variations, and restricts diffusion to mixtures involving a single diffusing pair without multi-component interactions. The basic model further excludes chemical reactions, focusing solely on passive scalar . These assumptions impose significant limitations in practical applications. In high-speed flows, introduces variable properties, rendering the constant Lewis number invalid as \alpha and D vary with and . For multi-component mixtures, the binary diffusion approximation fails to capture cross-species interactions, necessitating an effective diffusivity D to approximate the Lewis number. In reactive systems, such as , the Lewis number must be coupled with reaction timescales via the Damköhler number to account for reaction- interactions, which the standard form overlooks. Error sources further undermine accuracy, including the inherent temperature dependence of , where D \propto T^{3/2} per Chapman-Enskog theory, causing \mathrm{Le} to vary spatially and temporally in non-isothermal flows. The standard Lewis number also neglects Soret (thermal ) and Dufour (diffusion-thermo) effects, which induce cross-transport of and and become relevant in mixtures with significant temperature gradients. Validity is strongest in gases, where \mathrm{Le} \approx 1 reflects comparable and rates, but diminishes in liquids where \mathrm{Le} \gg 1 due to much slower , allowing to dominate over diffusive processes. For turbulent flows, extensions incorporate turbulent diffusivities \alpha_t and D_t to define an effective turbulent Lewis number \mathrm{Le}_t = \alpha_t / D_t, often assumed near unity for simplicity.

Applications

In Combustion Processes

In combustion processes, the Lewis number significantly influences and stability in premixed . For mixtures with Le < 1, such as lean hydrogen-air where Le ≈ 0.3–0.4, the faster of the deficient reactant relative to heat leads to an increased laminar and promotes cellular instabilities, resulting in wrinkled flame fronts that enhance overall propagation. Conversely, for Le > 1, as in rich mixtures, the slower mass stabilizes planar , reducing propagation speed and suppressing instabilities. In premixed flames, the Lewis number affects stretch sensitivity and extinction limits. Low Le flames exhibit reduced extinction under positive stretch due to enhanced reactant supply to curved regions, allowing survival at higher stretch rates compared to unity Le cases. For Le > 1, flames are more prone to extinction at lower stretch rates because heat diffuses faster than fuel, leading to local cooling in stretched zones. The Zeldovich-Frank-Kamenetskii theory assumes unity Le to simplify the analysis of flame propagation, equating thermal and mass diffusion profiles and yielding a straightforward expression for burning velocity without differential diffusion effects. In gaseous mixtures, Le is given by the ratio of the to the , typically near unity for many hydrocarbons but deviating in diluted or blended fuels. Diffusive-thermal instability in flames is critically dependent on the Lewis number and the density expansion ratio across the flame. Low Le (< 1) promotes this by causing preferential diffusion that depletes reactants in troughs and enriches peaks of perturbed flames, amplifying perturbations and leading to cellular structures, particularly near extinction limits. The instability criterion involves Le combined with the expansion ratio; for Le sufficiently below 1, it couples with hydrodynamic effects like Darrieus-Landau instability to destabilize planar fronts, with marginal stability occurring at specific wavenumbers dependent on Le (e.g., k ≈ 2 for Le = 0.4). In turbulent combustion, the Lewis number modulates scalar dissipation rates, which quantify the local mixing of fuel and oxidizer. For non-unity Le, particularly Le < 1, the transport equation for scalar dissipation shows enhanced dilatation effects and countergradient fluxes, acting as a sink that promotes mixing in distributed reaction zones by increasing gradient alignment with extensive strain. This leads to greater flame wrinkling and higher turbulent burning velocities in low-Le cases, such as hydrogen-enriched mixtures. Representative examples include methane-air flames at Le ≈ 1, which exhibit moderate stability with limited wrinkling under turbulence, while rich mixtures at Le > 1, like propane-air, suppress cellular structures and maintain near-planar propagation.

In Heat and Mass Transfer

In non-reactive processes, the Lewis number plays a crucial role in characterizing the relative rates of and mass diffusion, particularly in scenarios involving simultaneous transport such as and . During , the Lewis number determines the relative thicknesses of the (concentration) and the ; when Le ≈ 1, these layers are comparable, leading to coupled and mass transfer effects that influence the overall process efficiency. In water-air systems, typical values of Le ≈ 0.9 indicate nearly similar s, simplifying the modeling of rates in applications like cooling towers. For , a similar interplay occurs, where deviations from Le = 1 can alter the film formation and rejection characteristics. In drying processes, the affects the balance between conduction and within porous media, impacting removal . When Le < 1, diffusivity exceeds thermal diffusivity, accelerating vapor transport relative to propagation and potentially enhancing drying rates in the initial stages by reducing internal temperature gradients. This effect is particularly relevant in hygroscopic materials like wood or ceramics, where low Le values promote faster surface evaporation but can lead to uneven drying if supply lags, influencing energy consumption and product quality in industrial dryers. Conversely, higher Le values may slow transfer, requiring optimized airflow to maintain in porous structures. The analogy between convective heat and mass transfer is extended through the Lewis number in the Chilton-Colburn framework, providing a basis for predicting transfer coefficients in flowing fluids. Specifically, the relation between the heat transfer j-factor (j_H) and mass transfer j-factor (j_M) incorporates Le as j_H / j_M \approx \mathrm{Le}^{2/3}, allowing estimation of mass transfer rates from known heat transfer correlations when Prandtl and Schmidt numbers differ. This adjustment is essential for turbulent flows in ducts or over surfaces, where the Lewis number bridges the thermal and solutal boundary layers to yield accurate analogies without direct experimentation. In multiphase flows at gas-liquid interfaces, the Lewis number governs absorption rates by quantifying the disparity between thermal and mass diffusion in the liquid phase. For processes like gas absorption, low Le values enhance interfacial mass transfer by thinning the concentration boundary layer relative to the thermal one, increasing solute uptake. In CO_2 capture applications using aqueous absorbents, Le ≈ 1 in certain gas-phase dominated regimes simplifies predictive models, as heat and mass effects align closely, reducing computational complexity in reactor design. Industrial applications leverage the Lewis number to optimize heat and mass transfer in systems handling humid air or saline solutions. In heat exchangers processing humid air, such as air conditioning coils, Le is used to predict combined sensible and latent transfer coefficients, with values around 1 enabling simplified effectiveness-NTU methods for dehumidification performance. For desalination via humidification-dehumidification or membrane distillation, Le >> 1 for salt diffusion (due to low ionic diffusivity) highlights dominant heat transfer, guiding the design of evaporators to minimize scaling and maximize water flux. These insights ensure efficient operation by tailoring flow conditions to the specific Le of the working fluids.

In Biological Systems

In biological systems, the Lewis number (Le) plays a key role in governing the relative rates of and during processes like and , influencing efficiency and thermal balance in living organisms. For mammals, the high Le in aqueous environments (approximately 70–100 for oxygen diffusion in , due to being much greater than ) poses significant challenges for gill-like structures. If mammals relied on gills in , the thinner relative to the would result in rapid loss exceeding oxygen uptake, making such structures inefficient for maintaining and explaining the evolutionary preference for lungs in air-breathing . In plant transpiration, the Lewis number in air is close to unity (Le ≈ 0.85–1.0 for ), which facilitates balanced of and across boundary layers. This near-equality allows stomata to regulate effectively without disproportionate thermal gradients, supporting optimal cooling and CO₂ uptake under varying environmental conditions. For instance, in scenarios around leaves, the unity Le simplifies modeling of energy and mass balances, ensuring that evaporative losses align with transfer for sustained . At the cellular level, in biofilms and tissues, the Lewis number influences the interplay between /oxygen diffusion and heat propagation, often with high Le values in cytoplasmic environments (due to hindered mass diffusion from ). This promotes mass-limited , where oxygen or gradients develop faster than thermal ones, constraining reaction rates in dense microbial communities or matrices and favoring localized metabolic hotspots. The high Le in aqueous media has evolutionary implications, driving the transition from gill-based aquatic respiration to lung-based aerial breathing in terrestrial vertebrates. Gills, optimized for (where Le >> 1 leads to coupled but heat-dominant transfer), become maladaptive on land due to desiccation risks, whereas lungs exploit air's Le ≈ 1 for efficient, low-heat-loss . exemplify adaptation to water's high Le, with lamellar structures minimizing thermal mismatches to sustain ectothermy without excessive cooling. In , the Lewis relation (derived from Le ≈ 0.85 for in air) quantifies evaporative cooling rates, linking of sweat to convective heat loss. This enables precise , where skin wettedness and ambient gradients determine cooling efficiency, preventing overheating during exertion.

History

Origin and Naming

The is named after (1882–1975), a pioneering and the first head of the Department of at the , who made significant contributions to heat and mass transfer analogies during the . 's work laid foundational principles for modern by integrating physics, chemistry, and engineering to address industrial processes. The concept of the Lewis number first appeared implicitly in Lewis's 1922 paper, where he formalized the ratio of thermal diffusivity to mass diffusivity as part of analogies between heat and mass transfer mechanisms in evaporation processes. This key publication, titled "The Evaporation of a Liquid into a Gas," explored the mechanisms of liquid evaporation into gases, with applications to humidifiers, scrubbers, and related equipment, establishing the dimensionless ratio's role in quantifying transport similarities. Although the term "Lewis number" was not explicitly coined until later—gaining prominence in mid-20th-century combustion and transport literature, such as in the 1950s references to the analogy—the underlying formulation originated here, linking thermal conductivity, specific heat, and molecular diffusion in a rigorous manner. It is distinct from any association with Bernard Lewis (1899–1993), the renowned combustion scientist known for his work on flame propagation and explosives, despite occasional confusion in combustion-related literature due to shared naming and field overlaps. The Lewis number emerged within early 20th-century chemical engineering, particularly in the analysis of and columns, where analogies between and enabled simplified modeling of multicomponent separation processes. Lewis's contributions in this era, including his role in developing unit operations curricula, provided the practical context for the number's adoption in .

Development in Scientific Literature

In the 1930s and 1950s, the Lewis number gained prominence in combustion theory through its integration into models of flame propagation and structure, building on the thermal theory proposed by Mallard and Le Chatelier in 1883, which emphasized heat conduction ahead of the flame front and can be retroactively linked to Le effects on relative diffusion rates. This period saw formalization in Burke and Schumann's 1928 analysis of diffusion flames in cylindrical geometries, where a unity Lewis number assumption simplified the coupling of heat and mass transfer to predict flame shapes and positions. During the and , the Lewis number expanded into numerical simulations and (CFD) for turbulent reacting flows, with key contributions from D. B. Spalding's models that often approximated unity Le to couple scalar in premixed and diffusion flames. Influential texts like Bird, Stewart, and Lightfoot's (1960) standardized the Le as a fundamental parameter in multicomponent , enabling its widespread use in analyses of reacting systems. Later combustion references, such as Turns' An Introduction to (2012), detailed Le's role in flame stability and extinction limits, emphasizing its impact on preferential diffusion in non-unity cases. From the 1990s onward, the Lewis number found applications in microscale and , particularly in modeling and during where Le ≈ 1 governs diffusion-limited burning rates. NASA research during this era highlighted variable Le effects in hypersonic flows, with corrections to Le improving heat-transfer predictions for high-speed reacting boundary layers. By the 2020s, studies extended Le to plasma-assisted , where non-unity values influence ignition in NH₃/H₂/air mixtures under discharges. Recent advances as of 2025 incorporate models to predict in premixed , for example, a 2023 probabilistic approach for reaction rates in turbulent premixed under unity Lewis number assumptions, with plans to address non-unity effects on burning velocities without explicit transport equations.

References

  1. [1]
    What is Lewis Number - Definition - Thermal Engineering
    May 22, 2019 · The Lewis number is a dimensionless number, named after Warren K. Lewis (1882–1975). The Lewis number is defined as the ratio of thermal ...Missing: history | Show results with:history
  2. [2]
    Lewis number - tec-science
    May 10, 2020 · The Lewis number is a dimensionless similarity parameter to describe heat and mass transport. The Lewis number always comes into play when a flowing fluid is ...
  3. [3]
    Lewis Number - an overview | ScienceDirect Topics
    The Lewis number for air is approximately 1 (see Example 6), so with good ... (Warren Lewis has been called the father of US chemical engineering. He ...Missing: history | Show results with:history
  4. [4]
    Lewis Number - an overview | ScienceDirect Topics
    The Lewis number puts in correlation the mass diffusion and the thermal conductivity of a fluid. Similar to the Prandtl number, which correlates momentum ...
  5. [5]
    Lewis number - Goodwind
    Dec 1, 2024 · By definition, the Lewis number is the ratio of thermal diffusivity to mass diffusivity, expressed as:
  6. [6]
    Dimensionless numbers of the boundary layers (Prandtl, Schmidt ...
    May 15, 2020 · The Prandtl, Schmidt, and Lewis numbers are dimensionless numbers used to compare transport in boundary layers, relating two parameters, and ...Missing: 1 | Show results with:1
  7. [7]
    Effects of fuel Lewis number on the minimum ignition energy and its ...
    Jul 19, 2023 · ... fuel. The fuel Lewis number of 1.0 is representative of the stoichiometric methane–air mixture, whereas fuel Lewis numbers of 0.75 and 1.25 ...
  8. [8]
    What means Lewis number ? | ResearchGate
    Dec 25, 2018 · for example Le is around of 1 for gases flow and Le>>1 for liquids flow. It seems to me that this discussion is complete, unless there is some ...Missing: typical | Show results with:typical
  9. [9]
    Effects of Lewis number on turbulent kinetic energy transport in ...
    Jul 27, 2011 · The Lewis number is defined as the ratio of thermal diffusivity to mass diffusivity (i.e., Le = α T / D ⁠). Recent studies16–18 have indicated ...
  10. [10]
    (PDF) Experimental and numerical determination of Lewis number ...
    Apr 14, 2020 · The most appropriate effective Lewis number formulation is identified through comparison with experimentally extracted Lewis numbers (Le).
  11. [11]
    [PDF] _Chapter 4 - POLLUTANT EMISSION RATE
    The ratio of the Schmidt and Prandtl numbers is called the Lewis number (Le),. P ja. Sc k. Le. Pr c. = = ρ D. (4-77). We also introduce the Sherwood number (Sh) ...
  12. [12]
    [PDF] Role of Turbulent Prandtl Number on Heat Flux at Hypersonic Mach ...
    The turbulent Prandtl number is defined as t t t. /. Pr = (11). The choice of t α merits further elaboration. It was indicated in Ref. 4 that experiments in ...<|separator|>
  13. [13]
    [PDF] Lecture 9 Laminar Diffusion Flame Configurations
    Schmidt number is defined as Sc = ν/D. 9.-19. Page 20. The dimensionality of the problem may be reduced by introducing the similarity transformation which ...
  14. [14]
    [PDF] LA-7557-MS Simplified Multicomponent Phase Transition ... - OSTI
    system with a unit Lewis number for all species. The method will be ... B. R. Bird, W. E. Stewart, and E. N. Lightfoot, "Transport Phenomena,". (John ...<|control11|><|separator|>
  15. [15]
    [PDF] / N95- 21036 - NASA Technical Reports Server
    The problem simplifies if we can assume that the Lewis numbers of the species, i.e. the ratios of their thermal to their mass diffusivities, are equal to unity.
  16. [16]
    [PDF] Unsteady Strained Flames: Fundamentals and Numerical Modeling
    For a unity Lewis number, mass and heat diffusion rates are equal, and the reaction-diffusion zone is essentially unaffected over a range of weak to ...
  17. [17]
    Assessment of the constant non-unity Lewis number assumption in ...
    The current analysis suggests that, for numerous combustion configurations, the constant non-unity Lewis number approximation leads to small errors when the set ...
  18. [18]
    [PDF] Dimensional Analysis in Mass Transfer
    Jan 5, 2021 · Now, do the experiments. Peclet number. Pe ≡. Prandtl number. Pr ... between heat and mass. 1. Constant physical properties. 2. Small net ...
  19. [19]
    Criterion analysis and experimental study of combustion ...
    Nov 9, 2020 · In this case, the Lewis-Semenov number may differ from 1, and the similarity of the temperature distribution and concentration is violated.
  20. [20]
    Mass Ratio - an overview | ScienceDirect Topics
    For gases, it is Le∈⟨0.8; 1.2⟩, for fluids Le∈⟨70; 100⟩. In expression (2), it is usually called the Lewis–Semenov number, or sometimes only the Lewis number.
  21. [21]
    Thermosolutal Marangoni instability in a viscoelastic liquid film
    Dec 22, 2020 · The (inverse) Lewis number compares the characteristic mass diffusion time scale ({H^2}/D) with the thermal diffusion time scale ({H^2}/\alpha ) ...
  22. [22]
    [PDF] Effects of Lewis Number on Temperatures of Spherical Diffusion ...
    The numerical results show that the ambient gas Lewis number would have a strong effect on flame temperature if the flames were steady and nonradiating. For ...
  23. [23]
    Lewis number, Le, for saturated air and dry air at one atmosphere....
    Lewis number Le = D/α and thermal diffusivity α = k/ρC p. Numerical values of Le for dry air and saturated air at different temperatures are tabulated in Table ...
  24. [24]
    [PDF] an introduction to combustion: concepts and applications, third edition
    ... AN INTRODUCTION TO COMBUSTION. Concepts and Applications. THIRD EDITION. Stephen R. Turns ... Non-road. Miscellaneous. Figure 1.4. Trends in directly emitted ...
  25. [25]
    None
    ### Summary of Thermal and Concentration Boundary Layer Thickness and Lewis Number Relation
  26. [26]
    [PDF] 19740017703.pdf - NASA Technical Reports Server (NTRS)
    The second binary diffusion approximation is that the Lewis number is some constant value. Both approximations are used in this study. The multicomponent ...
  27. [27]
    [PDF] Effects of non-unity Lewis numbers in diffusion flames
    Non-unity Lewis numbers cause flame shifts toward fuel or oxygen, and the maximum temperature may not coincide with the flame.Missing: inverse | Show results with:inverse
  28. [28]
    Effects of thermophoresis, Soret-Dufour on heat and mass transfer ...
    Apr 15, 2020 · The effect of both Soret and Dufour are mostly neglected in the past due to their smaller order of magnitude as presented by Fick's laws. Soret ...
  29. [29]
    [PDF] Thermal diffusion coefficient modeling for high pressure combustion ...
    For example, the Lewis number in gases is usually of the order unity, which indicates that there is a comparable rate of heat and mass diffusion. In contrast, ...
  30. [30]
    Effective Lewis number and burning speed for flames propagating in ...
    The Lewis number dependence of the turbulent burning speed (31) and the turbulent Lewis number (33) appear to be in better agreement with the predictions of ...Missing: alpha_t D_t
  31. [31]
    Cellular instability in Le < 1 turbulent expanding flames
    In this investigation, we have examined the role of cellular instability on the propagation of turbulent flames for Le< 1, which inherently promotes DT ...
  32. [32]
    [PDF] Lewis Number Effects in Distributed Flames - CCSE
    The Lewis number, ratio of heat to mass diffusivity, affects flame response. High Lewis number flames may decrease burning rate, while low Lewis number flames ...Missing: liquids | Show results with:liquids
  33. [33]
    Effect of Lewis number on premixed laminar lean-limit flames ...
    It was shown that lean Le > 1 flames lose heat and extinguish at lower stretch rates while Le ≤ 1 flames are able to survive higher stretch rates. Along with ...
  34. [34]
    [PDF] Lecture 5 The Thermal Flame Theory
    Zeldovich and Frank-Kamenetzki equalize the derivatives of the preheat zone, and the reaction zone,. This yields an equation for the burning velocity. 5.-13 ...
  35. [35]
    [PDF] Diffusional-thermal instability of diffusion flames - Paul D. Ronney
    The patterns are observed experimentally for conditions under which the Lewis number of the fuel is less than unity, that is, the diffusion coefficient of the ...
  36. [36]
    Effects of Lewis Number on Scalar Dissipation Transport and Its ...
    The effects of Lewis number (Le) on scalar dissipation rate transport of the reaction progress variable have been studied using 3-dimensional DNS data of freely ...Missing: validity | Show results with:validity
  37. [37]
    Water Evaporation and Condensation in Air With Radiation: The Self ...
    This fact might be used to justify reduced-accuracy approximations in applying Spalding theory to air/water, such as Lewis number unity or Couette flow ( ...<|control11|><|separator|>
  38. [38]
    [PDF] Heat Transfer Workshop 11 - Water Evaporation Introduction Name
    For water vapor in air at typical room temperature conditions the Lewis number is nearly a constant value of 0.85. LLLL = 𝛼𝛼. DD. = 0.85. This also provides ...<|control11|><|separator|>
  39. [39]
    Lewis number in the context of air-drying of hygroscopic materials
    Lewis number (Le) is defined as the ratio of the thermal diffusivity to the mass diffusivity. It indicates if the heat and mass transfer processes are whether ...
  40. [40]
    Mass and Heat Transport Models for Analysis of the Drying Process ...
    May 16, 2018 · The modeling and numerical simulation of drying in porous media is discussed in this work by revisiting the different models of moisture migration.
  41. [41]
    [PDF] A numerical investigation of phase change effects in porous materials
    For a Lewis number greater than one, the opposite effect was observed. Another important result which can be observed in Fig. 5 is that the liquid ...
  42. [42]
    The analogy between heat and mass transfer in low temperature ...
    A modified version of the Chilton-Colburn analogy incorporating the Lewis Number of Evaporation was developed leading to a coefficient of determination of 0.96.
  43. [43]
    Hydrodynamics and gas-liquid mass transfer of CO2 absorption into ...
    The heat and mass transfer inside of a monolith honeycomb indicates that the Lewis number can deviate significantly from one in the entrance length [17].
  44. [44]
    The Behavior of Lewis Number in Finned Tube Cooling Coils under ...
    Aug 6, 2025 · It was found that the Lewis number (Le) varied within the range of 0.92-1.62 and that the increase in inlet air relative humidity tends to ...
  45. [45]
    (PDF) Application of Lewis analogy to estimate of the heat and mass ...
    Aug 10, 2025 · The heat transfer coefficient and the vapor-phase mass transfer coefficient decreased significantly with decreasing mass velocity. The mass ...
  46. [46]
    A Physical Modeling Approach for Higher Plant Growth in Reduced ...
    Indeed, the ratio between mass and heat boundary layer thickness is equal to Le1/3 where Le is the Lewis number, which is equal to unity in the case of gases ( ...
  47. [47]
    Structure, function and evolution of the gas exchangers - PMC - NIH
    Gills (evaginated gas exchangers) are the primordial respiratory organs: they are the archetypal water breathing organs. Lungs (invaginated gas exchangers) are ...
  48. [48]
    Heat Exchange Between Human Skin Surface and Thermal ...
    Jan 1, 2011 · 7.3 The Lewis Relation Between Heat and Mass ... Convective mass transfer and the coefficient of evaporative heat loss from human skin.
  49. [49]
    Warren Kendall Lewis | Biographical Memoirs: Volume 70
    Lewis published thirteen papers on distillation and nine on evaporation; nineteen of his eighty-one patents were on distillation. ... Warren K. Lewis was elected ...
  50. [50]
    Arthur D. Little, William H. Walker, and Warren K. Lewis
    Arthur D. Little, William H. Walker, and Warren K. Lewis were among the leaders of the movement to create the new profession of chemical engineering.Missing: 1922 | Show results with:1922
  51. [51]
    The Evaporation of a Liquid Into a Gas | J. Fluids Eng.
    Dec 5, 2023 · The author investigates the mechanism of the evaporation of a liquid into a gas as applied to such processes as are found in gas scrubbers, humidifiers, ...
  52. [52]
    The Lewis factor and its influence on the performance prediction of ...
    The relation of the Lewis factor to the Lewis number is also investigated. ... W.K. Lewis. The evaporation of a liquid into a gas. Trans. ASME, 44 (1922), pp ...
  53. [53]
    Bernard Lewis Gold Medal - The Combustion Institute
    The Bernard Lewis Gold Medal recognizes brilliant research in the field of combustion. The Gold Medal is presented biennially to one scientist.Missing: expert | Show results with:expert
  54. [54]
    [PDF] When Chemical Reactors Were Admitted And Earlier Roots of ...
    The success of Walker's program was assured when he arranged for a 1905 graduate, Warren Kendall Lewis, to go to Germany for a Ph. ... (1922), Lewis and Radasch's ...
  55. [55]
    The Industrial Relations of Science: Chemical Engineering at MIT
    For an example of a course of study organized around unit operations, see William H. Walker, Warren K. Lewis, and William H. McAdams, Principles of Chemical ...
  56. [56]
    Theory of flame propagation - ScienceDirect.com
    In Mallard and Le Chatelier's treatment of flame propagation the problem is considered simply one of heat flow in which the unburnt gas is raised to its ...
  57. [57]
    Diffusion Flames | Industrial & Engineering Chemistry
    Article October 1, 1928. Diffusion Flames. Click to copy article linkArticle link copied! S. P. Burke · T. E. W. Schumann. ACS Legacy Archive. Open PDF ...Missing: original | Show results with:original
  58. [58]
    Spherically Symmetric Droplet Combustion - Princeton University
    Apr 18, 1997 · For the combustion of most liquid fuels burning in air, the Spalding transfer number is typically between 1 and 10. Equations 1 through 3 ...
  59. [59]
    (PDF) Metal particle combustion and nanotechnology - ResearchGate
    Aug 9, 2025 · ... combustion is diffu-. sion-controlled. For diffusion control and a Lewis number of. unity (Le =a/D), the mass consumption rate of. a particle ...
  60. [60]
    Improved Heat-Transfer Calculations for Hypersonic Flow
    Jan 1, 1987 · Lewis number corrected for extremely high airspeeds. Algorithm calculates improved, variable value of Lewis number, factor in equation for ...
  61. [61]
    Plasma assisted NH3/H2/air ignition in nanosecond discharges with ...
    [14] experimentally investigated the influence of the Lewis number, pressure, and spark gap on the flow-facilitated ignition (FFI) phenomenon in electrode-spark ...
  62. [62]
    Probabilistic deep learning of turbulent premixed combustion
    Aug 8, 2023 · To accurately model non-unity Lewis number flames, an additional transport equation for the mixture fraction is needed.6–8. The filtering ...