Turbulent Prandtl number
The turbulent Prandtl number, denoted as \Pr_t, is a dimensionless quantity in fluid dynamics that characterizes the relative rates of turbulent momentum and heat transport in a fluid flow. It is defined as the ratio of the eddy viscosity \nu_t (or momentum diffusivity K_m) to the eddy thermal diffusivity \alpha_t (or heat diffusivity K_h), expressed mathematically as \Pr_t = \nu_t / \alpha_t = K_m / K_h.[1][2] This parameter plays a crucial role in turbulence modeling, particularly within Reynolds-Averaged Navier-Stokes (RANS) frameworks and first-order closure schemes, where it links the analogous transport of momentum and scalars like heat or mass in turbulent boundary layers.[1] In applications such as atmospheric dispersion, weather and climate simulations, and hypersonic aerothermodynamics, \Pr_t influences predictions of heat flux, shear stress, and scalar mixing, with dynamic models improving accuracy over constant-value assumptions.[1][2] Typical values of \Pr_t range from 0.7 to 1.0 in neutrally stratified turbulent flows, with a commonly adopted asymptotic value of 0.85 in many engineering contexts, though it can exceed 1.3 near walls or decrease under unstable stratification.[1][3] Variations arise from flow conditions, such as stability in the atmospheric boundary layer or low molecular Prandtl numbers in liquid metals, where specialized formulations like square-additive models adjust \Pr_t based on local turbulence variables.[4] Theoretically, \Pr_t extends the molecular Prandtl number concept to turbulent regimes, rooted in Boussinesq's eddy viscosity hypothesis and Monin-Obukhov similarity theory, enabling better parameterization of near-wall effects and scalar variance in computational fluid dynamics simulations.[1] In hypersonic flows, for instance, computing \Pr_t from two-equation models based on enthalpy variance significantly reduces discrepancies between theoretical heat flux predictions and experimental data.[2]Fundamentals
Definition
The turbulent Prandtl number, denoted Pr_t, is a dimensionless quantity that characterizes the relative rates of turbulent transport of momentum and heat in fluid flows. It is defined as the ratio of the eddy diffusivity for momentum \epsilon_m (also known as eddy viscosity) to the eddy diffusivity for heat \epsilon_h, expressed as Pr_t = \frac{\epsilon_m}{\epsilon_h}. [1][5] This parameter physically represents the ratio of the turbulent diffusivity of momentum to the turbulent diffusivity of heat, quantifying the relative efficiency of turbulence in transporting momentum versus scalar quantities like temperature when molecular diffusion effects are negligible.[1] In the gradient diffusion approximation commonly used in Reynolds-averaged Navier-Stokes modeling, the turbulent heat flux q_t is related to the mean temperature gradient via q_t = -\rho c_p \epsilon_h \frac{\partial T}{\partial y}, where \rho is the fluid density, c_p is the specific heat at constant pressure, and y is the coordinate normal to the flux direction; Pr_t then connects \epsilon_h to \epsilon_m for consistent transport modeling.[1] Unlike the molecular Prandtl number in laminar flows, which compares kinematic viscosity to thermal diffusivity at the molecular scale, Pr_t applies to eddy-mediated transport in turbulent regimes and typically takes values around 0.9 for many engineering and atmospheric flows, though it varies with factors such as flow stability and geometry.[6][1] The Reynolds analogy serves as a historical precursor, assuming Pr_t \approx 1 to equate momentum and heat transfer directly.[6]Relation to Laminar Flow
The laminar Prandtl number, denoted as Pr, is defined as the ratio of the kinematic viscosity \nu to the thermal diffusivity \alpha, given by Pr = \frac{\nu}{\alpha}. This dimensionless number characterizes the relative thickness of the momentum and thermal boundary layers in laminar flows and depends solely on the molecular properties of the fluid.[7] For common fluids, typical values include Pr \approx 0.7 for air at room temperature and Pr \approx 7 for water.[1][8] In contrast, the turbulent Prandtl number Pr_t emerges in turbulent flows and is determined by the ratio of eddy diffusivity for momentum to eddy diffusivity for heat, making it largely independent of the fluid's molecular properties and more sensitive to flow conditions such as stability and stratification.[7] While the laminar Pr remains a fixed fluid property, Pr_t typically ranges from 0.7 to 0.9 across various fluids in high-Reynolds-number regimes, reflecting the dominance of turbulent mixing over molecular diffusion. Historically, the concept of Pr_t developed alongside early turbulence theories, with foundational ideas from Taylor's 1915 work on eddy motion in the atmosphere and Boussinesq's eddy viscosity hypothesis, evolving to account for buoyancy effects by the 1940s.[7] As the flow Reynolds number increases beyond critical values (typically Re > 2300 for pipe flows), the transition from laminar to turbulent regime occurs, where molecular transport yields to turbulent eddies, rendering Pr_t the primary parameter for heat and momentum transfer in fully developed turbulence.[9] This distinction enables extensions of laminar flow analogies to turbulent cases, such as the Reynolds analogy assuming Pr_t near unity, with modifications like the Colburn analogy adjusting for molecular Pr to relate skin friction and heat transfer coefficients in turbulent boundary layers.[7]Theoretical Framework
Derivation from Turbulence Models
In the Reynolds-Averaged Navier-Stokes (RANS) framework, the turbulent Prandtl number arises from the modeling of turbulent stresses and fluxes through closure approximations. The Boussinesq hypothesis introduces an eddy viscosity \nu_t to relate the Reynolds stresses -\overline{u_i' u_j'} to the mean velocity gradients, analogous to the molecular viscosity in laminar flows: -\overline{u_i' u_j'} = \nu_t \left( \frac{\partial \overline{u_i}}{\partial x_j} + \frac{\partial \overline{u_j}}{\partial x_i} \right) - \frac{2}{3} k \delta_{ij}, where k is the turbulent kinetic energy and \delta_{ij} is the Kronecker delta.[10] This hypothesis simplifies the unknown Reynolds stress tensor by assuming a scalar eddy viscosity that scales the mean strain rate.[11] For heat transfer, a similar closure is applied to the turbulent heat flux -\overline{u_j' T'}, where T' is the fluctuating temperature. Under the gradient diffusion assumption, this flux is modeled as proportional to the mean temperature gradient: -\overline{u_j' T'} = \alpha_t \frac{\partial \overline{T}}{\partial x_j}, with \alpha_t denoting the turbulent thermal diffusivity.[10] The turbulent Prandtl number Pr_t then emerges as the ratio of the eddy viscosity to the eddy thermal diffusivity, Pr_t = \frac{\nu_t}{\alpha_t}, which quantifies the relative efficiency of turbulent momentum and heat transport. This relation parallels the laminar Prandtl number but accounts for the enhanced diffusivities in turbulent flows.[11] In the extension to two-equation turbulence models like the k-\epsilon model, Pr_t is incorporated into the transport equations for turbulence quantities. The eddy viscosity is expressed as \nu_t = C_\mu \frac{k^2}{\epsilon}, where C_\mu \approx 0.09 is a model constant, k is the turbulent kinetic energy, and \epsilon is its dissipation rate.[12] The k and \epsilon transport equations include diffusive terms modulated by turbulent Prandtl numbers \sigma_k and \sigma_\epsilon (typically \sigma_k = 1.0 and \sigma_\epsilon = 1.3), which govern the diffusion of k and \epsilon themselves: \frac{Dk}{Dt} = \frac{\partial}{\partial x_j} \left[ \left( \nu + \frac{\nu_t}{\sigma_k} \right) \frac{\partial k}{\partial x_j} \right] + P_k - \epsilon, \frac{D\epsilon}{Dt} = \frac{\partial}{\partial x_j} \left[ \left( \nu + \frac{\nu_t}{\sigma_\epsilon} \right) \frac{\partial \epsilon}{\partial x_j} \right] + C_{\epsilon 1} \frac{\epsilon}{k} P_k - C_{\epsilon 2} \frac{\epsilon^2}{k}, with production P_k and constants C_{\epsilon 1} = 1.44, C_{\epsilon 2} = 1.92. For the energy equation, Pr_t is often prescribed as a constant (e.g., 0.9) to close the model for scalar transport, ensuring consistency with the gradient diffusion hypothesis.[12] These constants, including the effective Pr_t, are calibrated from benchmark turbulent flows such as channel and boundary layer simulations.[11] The gradient diffusion hypothesis underpins the derivation of Pr_t by assuming that turbulent scalar fluxes follow a Fickian form, driven by local gradients much like molecular diffusion. This leads directly to Pr_t as the scaling factor between momentum and scalar eddy diffusivities, enabling the analogy between shear stress and heat flux expressions in near-wall regions.[10]Key Assumptions and Limitations
The turbulent Prandtl number, Pr_t, relies on the primary assumption of isotropic eddy diffusivities for momentum and heat, implying that turbulent transport mechanisms are directionally uniform and analogous across scalars. This assumption underpins first-order closure schemes in Reynolds-averaged Navier-Stokes (RANS) models, where eddy viscosity and diffusivity are treated similarly regardless of flow orientation. However, it fails in non-homogeneous or buoyant flows, where thermal stratification induces anisotropy in turbulent structures, leading to directional variations in eddy diffusivities that distort Pr_t predictions.[1][13] In flows with variable properties, such as those influenced by significant temperature or pressure gradients, the assumption of a constant Pr_t becomes invalid, as the ratio of eddy diffusivities shifts due to changes in fluid density, viscosity, and thermal conductivity. For instance, in combustion processes, Pr_t can vary substantially across regions of heat release, with simulation results showing deviations from typical constant values (around 0.9) ranging from 0.5 to 2.0, thereby invalidating fixed-Pr_t models and affecting mixing efficiency.[3] The Boussinesq eddy viscosity hypothesis, central to deriving Pr_t in turbulence models, further simplifies turbulent stresses by assuming they align isotropically with mean velocity gradients through a scalar eddy viscosity. This over-simplification neglects inherent anisotropy in turbulence, particularly in complex configurations like swirling or separated flows, where rotational effects or flow detachment generate non-uniform stress tensors, resulting in predictive errors for heat and momentum transport.[14][15] Advanced turbulence simulations highlight the need for variable Pr_t formulations, especially contrasting large eddy simulations (LES) with RANS approaches. In RANS, a constant Pr_t is often sufficient for averaged flows but underperforms in resolving local variations; conversely, LES requires a subgrid-scale Pr_t that accounts for unresolved anisotropic effects and filter-scale dependencies, enabling more accurate capture of buoyant or stratified dynamics without the isotropic constraints of RANS.[1][16]Applications
Heat Transfer in Turbulent Flows
In turbulent flows, the turbulent Prandtl number Pr_t plays a crucial role in predicting heat transfer rates through its incorporation into Nusselt number correlations, which relate convective heat transfer to flow parameters. The Nusselt number Nu is typically expressed as a function of the Reynolds number Re, the molecular Prandtl number Pr, and Pr_t, accounting for the relative efficiencies of turbulent momentum and heat transport. For instance, the widely used Dittus-Boelter correlation for fully developed turbulent flow in smooth pipes, Nu = 0.023 Re^{0.8} Pr^{0.4} (for heating), implicitly assumes Pr_t \approx 0.9 in its derivation based on eddy diffusivity models, enabling accurate estimation of heat transfer coefficients for fluids like air and water where Pr \approx 1. This assumption enhances the correlation's applicability in engineering designs, such as boiler tubes, by capturing the turbulent enhancement of thermal diffusion without explicit Pr_t specification.[17][18] In turbulent boundary layers, Pr_t directly influences the temperature profile, particularly in the logarithmic region where the mean temperature follows a modified log-law analogous to the velocity log-law. The temperature distribution is given by T^+ = Pr_t \left( u^+ + C_T \right), where T^+ and u^+ are dimensionless temperature and velocity, respectively, and C_T is an additive constant; deviations from Pr_t = 1 adjust the slope, reflecting differences in eddy diffusivities for heat and momentum. This modification is essential for modeling thermal boundary layers over flat plates or airfoils, as it predicts how heat flux varies with wall distance and ensures consistency with experimental profiles in high-Re flows. Seminal analyses confirm that Pr_t near 0.9 optimizes agreement between predicted and measured temperature gradients in the overlap region.[19][20] For convective heat exchangers, Pr_t affects the Stanton number St = \frac{Nu}{Re Pr}, which quantifies the ratio of heat transfer to fluid enthalpy flux and is critical for sizing exchanger surfaces. Turbulent mixing amplifies St beyond laminar predictions, with Pr_t calibrating the analogy between skin friction and heat transfer; lower Pr_t values enhance thermal penetration relative to momentum, increasing St by up to 20% in cross-flow configurations for gases. This adjustment is vital in compact heat exchanger designs, such as those in automotive radiators, where Pr_t-dependent correlations refine efficiency calculations under varying load conditions.[21] A notable application occurs in atmospheric boundary layers under neutral stability, where Pr_t \approx 0.74 governs sensible heat flux profiles, linking surface heating to vertical temperature gradients. This value, derived from Monin-Obukhov similarity theory, ensures accurate modeling of near-neutral conditions in weather forecasting and pollutant dispersion, as it balances turbulent diffusivities in the surface layer without buoyancy effects. Field observations over flat terrain validate this Pr_t, highlighting its role in predicting heat transfer from land to air in stable environments.[22][23]Mass Transfer and Momentum Transport
In turbulent flows, the turbulent Prandtl number (Pr_t) plays a central role in mass transfer through its analogy to the turbulent Schmidt number (Sc_t), which characterizes the ratio of momentum diffusivity to mass diffusivity in the turbulent regime. Specifically, Pr_t is defined as the ratio of the eddy diffusivity for momentum (ε_m) to the eddy diffusivity for heat (ε_h), while Sc_t is the analogous ratio ε_m / ε_c, where ε_c denotes the eddy diffusivity for mass concentration. For passive scalars, empirical and modeling studies often assume Pr_t ≈ Sc_t, implying that the turbulent transport of concentration fluxes mirrors that of momentum, with deviations arising from scalar-specific effects like buoyancy or molecular properties. This analogy facilitates the prediction of mass transfer rates using momentum transport correlations, particularly in high-Reynolds-number flows where molecular diffusion is negligible compared to turbulent mixing.[1][7] The turbulent Prandtl number governs the spreading rate of scalar plumes in mixing layers by determining the relative diffusion of scalar fields (such as concentration) compared to velocity profiles. In free shear flows like jets and plumes, a lower Pr_t (typically around 0.7) enhances scalar spreading relative to momentum, leading to broader concentration profiles and increased entrainment of ambient fluid into the plume. Direct numerical simulations of turbulent jets and plumes confirm that Pr_t influences the entrainment coefficient, which scales the plume's radial growth and mixing efficiency, with values derived from the balance between buoyancy and turbulent diffusion. This effect is critical in environmental dispersion models, where scalar plumes from point sources spread faster than momentum wakes when Pr_t < 1, altering plume dilution rates.[24][25] The linkage between mass transfer and momentum transport is most pronounced under the Reynolds analogy, which assumes Pr_t = 1, implying identical turbulent diffusivities for momentum, heat, and mass, and thus a direct proportionality between wall shear stress and scalar fluxes. This perfect analogy holds approximately in low-Prandtl-number drag-reducing flows, such as those with polymer additives, where it simplifies predictions of friction reduction and mass transfer enhancement by equating skin friction coefficients to Sherwood numbers. In such studies, deviations from Pr_t = 1 reveal selective suppression of momentum transport relative to scalar diffusion, enabling optimized designs for pipelines and heat exchangers.[26][27] In chemical reactors, the turbulent Prandtl number influences the thickness of reaction zones during turbulent combustion and dissolution processes by modulating the mixing of reactants across scalar gradients. For instance, in non-premixed flames, a Pr_t near 0.7–0.85 broadens the reaction zone compared to the momentum boundary layer, promoting more uniform reactant distribution and reducing hotspots, as evidenced in confined turbulent flame simulations where scalar diffusivity affects flame brush thickness. Similarly, in dissolution of solids within turbulent flows, higher Pr_t values confine the concentration boundary layer, slowing mass transfer rates and extending dissolution times, which is vital for reactor efficiency in pharmaceutical and petrochemical applications. These effects underscore Pr_t's role in scaling reaction rates with turbulence intensity, guiding computational models for safer and more productive reactor designs.[28][29]Determination and Values
Experimental Measurements
Experimental measurements of the turbulent Prandtl number, defined as the ratio of eddy diffusivity for momentum to that for heat (\Pr_t = \epsilon_m / \epsilon_h), have evolved from rudimentary probe-based techniques to advanced optical and computational methods. In the 1950s, pioneering efforts relied on pitot-static probes for mean velocity profiles and fine thermocouples for temperature measurements in turbulent boundary layers over flat plates and in pipes. These early setups, often conducted in wind tunnels or water channels, allowed estimation of \Pr_t by integrating mean profiles to infer turbulent transport coefficients, yielding values around 0.9 for air flows despite limitations in resolving fluctuations. Early works in the 1950s, building on boundary layer theories, established empirical estimates of \Pr_t \approx 1 for moderate molecular Prandtl numbers through profile integrations, highlighting its near-unity behavior.[30] Hot-wire anemometry emerged in the late 20th century as a primary technique for directly capturing simultaneous velocity and temperature fluctuations, enabling computation of \Pr_t = \frac{ -\overline{u'v'} }{ -\overline{v'T'} } \cdot \frac{ d\Theta/dy }{ dU/dy } from measured correlations and mean gradients. Single- and X-wire probes, operated in constant-temperature mode, measure fluctuating velocities, while cold-wire probes detect temperature variance; their cross-correlation provides the turbulent heat flux relative to Reynolds stress. For instance, in zero-pressure-gradient boundary layers over heated flat plates, such measurements at Reynolds numbers up to $10^5 revealed \Pr_t \approx 0.85 in the logarithmic region, with deviations near the wall due to conductive effects. This method has been widely applied in rotating and non-isothermal boundary layers, where rotation alters \Pr_t by influencing burst-sweep structures.[31][32] Laser Doppler velocimetry (LDV), often coupled with fine-wire thermocouples or optical thermometry, offers non-intrusive velocity measurements in pipes and channels, complemented by temperature data to estimate \Pr_t. In pipe flows with surfactant additives for drag reduction, two-component LDV tracks particle velocities while 1-μm thermocouples resolve temperature fluctuations, showing \Pr_t > 1 near walls due to suppressed turbulence. For fully non-intrusive approaches, LDV pairs with coherent anti-Stokes Raman scattering (CARS) or Rayleigh scattering in jets and channels, providing instantaneous velocity-temperature correlations at high Reynolds numbers (e.g., Re ≈ 21,000), where \Pr_t approaches 0.85 in the core. These techniques minimize probe interference, ideal for confined geometries like channels with heated walls.[33][34] Modern advancements include particle image velocimetry (PIV) and planar laser-induced fluorescence (PLIF) for planar fields in channels and jets, evolving from 1950s probe methods to high-resolution imaging. PIV captures velocity vectors via tracer particles, while PLIF maps scalar (temperature or concentration) distributions using fluorescent dyes, allowing \Pr_t computation from integrated fluxes. In stratified jets, simultaneous PIV/PLIF at Re ≈ 5,000 revealed spatial variations in \Pr_t from 0.7 to 1.2, influenced by buoyancy. These optical methods excel in complex flows, providing two-dimensional data for validation. Complementing experiments, direct numerical simulations (DNS) serve as benchmarks; for channel flow at friction Reynolds number Re_τ = 180 and molecular Pr = 0.71, DNS yields \Pr_t \approx 0.85 in the log layer, aligning with hot-wire data and confirming near-constancy away from walls.[35]Typical Values and Variations
In wall-bounded shear flows, such as those in pipes and boundary layers, the turbulent Prandtl number typically ranges from 0.85 to 0.9 under neutral conditions and high Reynolds numbers.[36] This value reflects the near-equality of eddy diffusivities for momentum and heat in fully developed turbulent regimes.[37] Significant variations occur across flow types. In free shear flows like jets and wakes, Pr_t is generally lower, around 0.7, due to enhanced scalar mixing relative to momentum transport. In buoyant or compressible flows, Pr_t can increase to 1.2–1.5 under stable stratification, where buoyancy suppresses vertical heat transport more than momentum, or decrease below 0.74 in unstable conditions as convective plumes enhance scalar dispersion.[38] Several factors influence these values. Near the wall (y⁺ < 10), Pr_t increases to approximately 1.1 owing to stronger suppression of heat transport relative to momentum, transitioning to molecular Pr in the viscous sublayer.[39][40] Stratification stability further modulates Pr_t, with lower values in unstable setups promoting scalar transport.[7] Reynolds number effects are weak for Re > 10⁴, where Pr_t stabilizes, though slight decreases may occur at lower Re due to incomplete turbulence development.[36] Empirical data from literature illustrate these ranges, as summarized below:| Flow Type | Pr_t Value | Conditions | Source |
|---|---|---|---|
| Fully developed duct flow | 0.92 | Neutral, high Re, air (Pr=0.7) | Kays and Crawford (1993) |
| Pipe flow (Poiseuille) | 0.8–0.9 | Channel center, Pr > 0.7 | Papavassiliou (2010)[39] |
| Atmospheric boundary layer | 0.73–0.92 | Neutral, laboratory | Li et al. (2018)[7] |
| Stable buoyant flow | >1.0 | Increasing Richardson number (Rg) | Katul et al. (2016)[38] |