Fact-checked by Grok 2 weeks ago

Turbulence modeling

Turbulence modeling is the process of developing mathematical approximations to represent the effects of in flows, enabling efficient simulations in (CFD) without resolving every turbulent fluctuation. These models are essential for predicting complex, chaotic behaviors in engineering applications such as , , and , where direct computation of all scales is computationally prohibitive. The field has evolved over the past 50 years alongside advances in computing power, beginning with early Reynolds-averaged approaches in the and progressing to more sophisticated methods today. Key challenges include balancing accuracy with computational cost, as spans a wide range of scales from large eddies to microscopic , often requiring empirical closures based on experimental data. Influential frameworks draw from statistical theories, such as those addressing the in isotropic . The primary categories of turbulence models include Reynolds-Averaged Navier-Stokes (RANS), which time-averages the Navier-Stokes equations to model all turbulent scales via additional transport equations for quantities like turbulent ; Large Eddy Simulation (LES), which explicitly resolves large-scale eddies while subgrid-scale models handle smaller ones; and (DES), a hybrid of RANS and LES for improved near-wall and separated flow predictions. Within RANS, widely used variants are the k-ε model, robust for free shear flows and employing equations for turbulent (k) and dissipation rate (ε); the k-ω model, superior near walls with equations for k and specific dissipation rate (ω); and the Shear Stress Transport (SST) model, blending k-ω and k-ε for versatile boundary layer simulations. These approaches continue to advance through integrations and high-fidelity validations, enhancing reliability in .

Fundamentals of Turbulent Flows

Characteristics of Turbulence

Turbulent flows represent a regime of fluid motion distinct from laminar flows, first systematically identified through pipe flow experiments conducted by Osborne Reynolds in 1883. In these experiments, Reynolds injected dye into water flowing through a glass tube and observed that at low flow velocities, the dye streak remained straight and coherent, indicative of smooth, layered (laminar) motion; however, beyond a critical velocity, the streak diffused rapidly into irregular, sinuous patterns, marking the onset of turbulence. This transition depends critically on the Reynolds number, a dimensionless parameter defined as the ratio of inertial to viscous forces, typically exceeding values around 2,000–4,000 in pipes for the shift to occur. Reynolds' work established that turbulence arises in flows where inertial effects dominate, leading to unpredictable, time-dependent behavior rather than steady, predictable paths. The hallmark of turbulence is its irregular, chaotic nature, characterized by the presence of eddies spanning a vast of scales, from large structures comparable to the domain down to microscopic sizes. This multi-scale vorticity-dominated motion is inherently three-dimensional, with fluctuations occurring in all spatial directions and lacking any preferred , unlike the predominantly two-dimensional streamlines of laminar flows. Turbulence emerges and persists only at sufficiently high Reynolds numbers, where viscous is insufficient to suppress instabilities, resulting in a self-sustaining cascade of among eddies. As articulated in foundational analyses, this irregularity manifests as random variations in , , and other quantities over time and , defying deterministic prediction without averaging or statistical treatment. A key feature of turbulence is the , theorized by in 1941, whereby injected at large scales by mean flow instabilities is transferred nonlinearly to progressively smaller eddies through vortex interactions, without significant viscous losses in the intermediate inertial subrange. In this subrange, the energy spectrum follows a universal -5/3 scaling with , independent of the fluid's or the large-scale forcing, until reaching the dissipation scale—known as the Kolmogorov microscale—where viscous effects dominate and is converted to heat. This process underscores turbulence's dissipative yet efficient nature. Additionally, turbulence exhibits , with intense local fluctuations in velocity gradients and occurring sporadically amid calmer regions, and vorticity amplification via and tilting mechanisms that intensify rotational motion and sustain the turbulent state across scales. These characteristics have profound practical implications in applications, where turbulence enhances molecular mixing, elevates on surfaces, and augments convective rates compared to laminar regimes. In , for instance, turbulent boundary layers delay , reducing overall drag on wings while increasing local heat loads during high-speed flight; in systems, the vigorous mixing promotes rapid fuel-air interactions, improving efficiency and reducing emissions in engines and burners. Such effects necessitate specialized modeling to predict and control turbulent phenomena accurately in design. Turbulent flows are typically analyzed statistically, with correlations capturing average behaviors amid the chaos. The Navier-Stokes equations constitute the fundamental set of partial differential equations governing the motion of viscous fluids, encompassing both laminar and turbulent regimes. For incompressible flows, where \rho is constant and the is low, the equations consist of the momentum balance \frac{\partial \mathbf{u}}{\partial t} + (\mathbf{u} \cdot \nabla) \mathbf{u} = -\frac{1}{\rho} \nabla p + \nu \nabla^2 \mathbf{u} + \mathbf{f}, coupled with the continuity equation \nabla \cdot \mathbf{u} = 0, where \mathbf{u} is the velocity vector, p is the , \nu is the kinematic , and \mathbf{f} represents body forces. In compressible flows, \rho varies, necessitating the full set including a \frac{\partial \rho}{\partial t} + \nabla \cdot (\rho \mathbf{u}) = 0, a momentum equation of similar form but with variable \rho and additional terms for viscous stresses, and an energy equation to close the system, allowing for phenomena like shock waves and variable temperature effects. The nonlinear advection term (\mathbf{u} \cdot \nabla) \mathbf{u} is the primary source of the and multi-scale observed in , as it couples different components and generates interactions across scales, leading to and in high-Reynolds-number flows. This nonlinearity amplifies small perturbations into large-scale structures while cascading energy to smaller scales, resulting in the irregular, three-dimensional fluctuations characteristic of turbulent motion. Direct Numerical Simulation (DNS) aims to solve the Navier-Stokes equations without approximations, fully resolving all turbulent scales from the largest eddies to the smallest dissipative structures. This requires spatial grid spacing on the order of the Kolmogorov length scale \eta = (\nu^3 / \epsilon)^{1/4}, where \epsilon is the dissipation rate, and time steps comparable to the Kolmogorov time scale \tau_\eta = (\nu / \epsilon)^{1/2}, to capture the dynamics accurately. The computational cost scales as O(\mathrm{Re}^{9/4}), where \mathrm{Re} is the , due to the increasing number of grid points and time steps needed as \mathrm{Re} grows, rendering DNS prohibitive for high-\mathrm{Re} flows typical in applications. DNS remains feasible primarily for low-to-moderate Reynolds number cases, such as turbulent channel flow at friction Reynolds numbers \mathrm{Re}_\tau \approx 2000, where full resolution is achievable with current computational resources. However, even at these levels, the simulations demand immense resources—on the order of billions of grid points and extended run times—and become impractical for industrial-scale problems involving complex geometries or \mathrm{Re} > 10^4, where alternative modeling approaches are essential.

Statistical Approaches to Turbulence

, characterized by its chaotic and irregular nature, is analyzed through statistical methods that quantify its probabilistic behavior rather than deterministic trajectories. These approaches rely on averaging procedures to extract mean properties and fluctuations from the multitude of possible realizations, enabling predictive modeling despite the inherent unpredictability. Ensemble averaging provides a fundamental statistical tool for , defined as the of a quantity f over an infinite number of realizations: \langle f \rangle = \lim_{N \to \infty} \frac{1}{N} \sum_{i=1}^N f_i, where each f_i represents the value in the i-th realization under identical initial and boundary conditions. This method captures the probabilistic essence of by considering the ensemble as a of all possible outcomes, particularly useful for non- or inhomogeneous flows where multiple experiments can be conducted. For practical turbulent flows, time averaging serves as an alternative, computed as \langle f \rangle = \lim_{T \to \infty} \frac{1}{T} \int_0^T f(t) \, dt, assuming the flow statistics do not evolve with time after an initial transient. This temporal approach is widely applied in contexts, such as experiments, where long-duration measurements approximate the ensemble mean under suitable conditions. A key outcome of these averaging techniques is the tensor, which quantifies the turbulent flux through the correlation of velocity fluctuations: R_{ij} = \langle u_i' u_j' \rangle, where u_i' and u_j' are the fluctuating components of the velocity field relative to the . This represents the additional stresses arising from turbulent eddies transporting , analogous to molecular stresses in laminar flows but orders of magnitude larger in high-Reynolds-number regimes. The diagonal elements R_{ii} (no sum) correspond to turbulent contributions, while off-diagonal terms like R_{12} indicate stresses that drive exchange across streamlines. Higher-order statistics extend this analysis to capture non-Gaussian features of , such as , where intense bursts of activity alternate with quiescent periods. , defined as S = \langle (u')^3 \rangle / \langle (u')^2 \rangle^{3/2}, measures the in fluctuation distributions, often negative in boundary layers due to sweep events. , or , F = \langle (u')^4 \rangle / \langle (u')^2 \rangle^2, quantifies the prevalence of extreme events; values exceeding 3 (the Gaussian limit) indicate intermittent structures, with typical turbulent flows showing F \approx 4-10 in the inertial range. These moments reveal deviations from , essential for understanding energy dissipation variability. Two-point correlations provide insights into spatial structures and scale interactions, forming the basis for spectral analysis in turbulence. The two-point velocity correlation function R_{ij}(\mathbf{r}) = \langle u_i(\mathbf{x}) u_j(\mathbf{x} + \mathbf{r}) \rangle describes how fluctuations at separated points are related, enabling the derivation of the energy spectrum E(k). In the inertial subrange of isotropic turbulence, Kolmogorov's theory predicts E(k) \propto \epsilon^{2/3} k^{-5/3}, where \epsilon is the dissipation rate and k the wavenumber, reflecting a universal cascade of energy from large to small scales independent of viscosity. In homogeneous turbulence, the ergodicity assumption equates ensemble and time averages, stating that a single long-time realization suffices to represent the full statistical ensemble due to the mixing nature of the flow. This hypothesis holds under spatial uniformity, allowing practical computations from time series in simulations or measurements, though it breaks down in inhomogeneous cases with persistent structures.

Reynolds-Averaged Navier-Stokes Framework

Reynolds Decomposition and Averaging

Reynolds decomposition is a foundational technique in turbulence analysis that separates the instantaneous field into a time-averaged mean component and a fluctuating component. Introduced by Osborne Reynolds in his 1895 paper on the dynamics of incompressible viscous fluids, this method addresses the irregular nature of turbulent flows by enabling statistical treatment of the fluctuations. For a component u(\mathbf{x}, t), the decomposition is expressed as u(\mathbf{x}, t) = \overline{u}(\mathbf{x}) + u'(\mathbf{x}, t), where \overline{u}(\mathbf{x}) denotes the mean , independent of time, and u'(\mathbf{x}, t) represents the fluctuating part, with its average defined to be zero: \langle u' \rangle = 0. This separation applies to all flow variables, such as and , facilitating the study of mean flow behavior while accounting for turbulent contributions. The averaging operation, typically a time average for turbulence, satisfies several key properties that underpin its utility in deriving statistical equations. holds, so \langle a + b \rangle = \langle a \rangle + \langle b \rangle and \langle c \cdot f \rangle = c \cdot \langle f \rangle for a constant c. Additionally, the average of a product decomposes as \langle ab \rangle = \langle a \rangle \langle b \rangle + \langle a' b' \rangle, highlighting the of fluctuations that drives turbulent . The averaging commutes with spatial and temporal derivatives, ensuring \langle \frac{\partial f}{\partial t} \rangle = \frac{\partial \langle f \rangle}{\partial t} and similarly for spatial derivatives, which preserves the structure of governing equations under averaging. These properties rely on fundamental assumptions about the flow. Reynolds averaging assumes statistical stationarity, meaning flow statistics are invariant over the , allowing a well-defined time for the . Homogeneity may also be invoked for spatial averaging in uniform directions, though it is not strictly required for the itself. For non-stationary flows with periodic components, such as those involving organized structures like wakes or jets, the standard is extended to a triple : u(\mathbf{x}, t) = \overline{u}(\mathbf{x}) + \tilde{u}(\mathbf{x}, t) + u''(\mathbf{x}, t), where \tilde{u} captures the periodic with zero over a cycle, and u'' denotes the random turbulent fluctuation, satisfying \langle u'' \rangle = 0 and \langle \tilde{u} \rangle = 0. This approach, developed by Hussain and Reynolds in to isolate coherent motions in turbulent shear flows, enables separate analysis of , organized, and contributions.

Derivation of RANS Equations

The derivation of the Reynolds-averaged Navier-Stokes (RANS) equations proceeds by applying the —introduced by Osborne Reynolds in his foundational work on —to the instantaneous incompressible Navier-Stokes equations, separating each flow variable into its time-averaged mean and fluctuating components. For the velocity field, this decomposition is expressed as u_i = \overline{u}_i + u_i', where \overline{u}_i denotes the mean velocity and u_i' the fluctuating part with zero mean (\langle u_i' \rangle = 0); a similar decomposition applies to as p = \overline{p} + p'. This approach, as detailed in standard turbulence texts, enables the extraction of equations governing the mean flow while accounting for turbulent fluctuations through statistical averaging. The instantaneous incompressible Navier-Stokes equations are \frac{\partial u_i}{\partial t} + u_j \frac{\partial u_i}{\partial x_j} = -\frac{1}{\rho} \frac{\partial p}{\partial x_i} + \nu \frac{\partial^2 u_i}{\partial x_j \partial x_j}, with the \frac{\partial u_i}{\partial x_i} = 0. Substituting the Reynolds decomposition into the continuity equation and taking the time average yields the continuity equation for the mean flow: \frac{\partial \overline{u}_i}{\partial x_i} = 0, since the average of the fluctuating velocity divergence is zero. This equation ensures mass conservation for the mean velocity field \overline{u}_i. For the momentum equation, substitution of the decomposition into the convective term u_j \frac{\partial u_i}{\partial x_j} produces (\overline{u}_j + u_j') \frac{\partial (\overline{u}_i + u_i')}{\partial x_j}. Upon time-averaging, the linear terms simplify to \overline{u}_j \frac{\partial \overline{u}_i}{\partial x_j}, while the nonlinear fluctuating contributions average to the additional term \frac{\partial \overline{u_i' u_j'}}{\partial x_j}. The pressure and viscous terms average directly to their mean counterparts, as fluctuations in pressure and the Laplacian of velocity average to zero in this formulation. The resulting RANS momentum equation is thus \frac{\partial \overline{u}_i}{\partial t} + \overline{u}_j \frac{\partial \overline{u}_i}{\partial x_j} = -\frac{1}{\rho} \frac{\partial \overline{p}}{\partial x_i} + \nu \frac{\partial^2 \overline{u}_i}{\partial x_j \partial x_j} - \frac{\partial \overline{u_i' u_j'}}{\partial x_j}. Here, the term -\frac{\partial \overline{u_i' u_j'}}{\partial x_j}, arising from the Reynolds stress correlations \overline{u_i' u_j'}, acts as an additional "turbulent diffusion" flux in the momentum equation, representing the transport of mean momentum by turbulent fluctuations. Analogous averaging applies to the energy and scalar transport equations derived from the Navier-Stokes framework. For instance, the mean temperature or \overline{\theta} satisfies an averaged advection-diffusion with an extra term -\frac{\partial \overline{\theta' u_j'}}{\partial x_j}, capturing turbulent scalar flux. These derived close the system for the mean flow only if the unknown correlations—such as the six independent components of the symmetric tensor \overline{u_i' u_j'}—are modeled, as they cannot be determined solely from the mean flow variables. This modeling requirement, known as the closure problem, necessitates additional relations beyond the RANS equations themselves.

Emergence of the Closure Problem

The Reynolds-Averaged Navier-Stokes (RANS) equations describe the evolution of mean flow quantities in turbulent flows but introduce additional unknown terms through the tensor, \overline{u_i' u_j'}, which represents the correlation of velocity fluctuations. For incompressible flows, the instantaneous Navier-Stokes equations consist of four equations ( and three momentum equations) for four unknowns (three velocity components and ). Upon Reynolds averaging, these become four equations for the mean quantities, but the Reynolds stresses add six independent components (due to symmetry of the tensor), resulting in 10 unknowns: three mean velocities, mean , and six Reynolds stress components. This underdeterminacy, known as the closure problem, renders the system mathematically incomplete without additional relations to express the Reynolds stresses in terms of the solvable mean fields. A deeper reveals the problem as part of an infinite hierarchy of in statistics. The exact transport for the second- Reynolds stresses \overline{u_i' u_j'} can be derived from the Navier-Stokes applied to the fluctuating fields, but it involves unclosed correlations \overline{u_i' u_j' u_k'}, which appear in the nonlinear and pressure-strain terms. Similarly, for these third moments introduce fourth-order correlations, leading to an unending regress where higher-order perpetually require further approximations. This hierarchy underscores the fundamental challenge in statistical theory: no finite set of is self-contained without or modeling assumptions. The need for was recognized early in research, with in 1925 highlighting the empirical necessity to relate turbulent stresses to mean gradients via concepts like mixing length, acknowledging that exact solutions were infeasible without such approximations. Later, Andrey Kolmogorov's 1941 theory of spectra provided a statistical framework emphasizing the cascade of energy across scales from large, energy-containing eddies to small, dissipative ones, further contextualizing the multi-scale nature of interactions that defy exact at any single level. These insights established that empirical or semi-empirical models are essential to approximate the unresolved scales. The implications of the problem are profound for turbulence modeling: any practical approach must approximate the non-local and anisotropic effects of —such as momentum redistribution across disparate scales—without explicitly resolving the full spectrum of fluctuations, balancing computational feasibility with physical fidelity in engineering applications like and .

Core Concepts in Turbulence Closure

Boussinesq Eddy Viscosity Hypothesis

The Boussinesq eddy viscosity hypothesis, first proposed by Joseph Valentin Boussinesq in 1872 and published in 1877, provides a foundational approach to modeling the Reynolds stresses in turbulent flows by drawing an analogy to molecular viscous stresses in laminar flows. In the Reynolds-averaged Navier-Stokes (RANS) framework, the Reynolds stresses -\overline{u_i' u_j'} represent the turbulent momentum flux, and Boussinesq postulated that these can be expressed in terms of the mean velocity gradients through an effective turbulent \nu_t: -\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 = \frac{1}{2} \overline{u_m' u_m'} is the turbulent kinetic energy and \delta_{ij} is the . This formulation assumes that turbulent eddies transfer momentum similarly to but at a much larger scale, with \nu_t \gg \nu (the molecular kinematic ) in high-Reynolds-number flows. The analogy to laminar flows is direct: in Newtonian fluids, the viscous stress tensor is \tau_{ij} = \mu \left( \frac{\partial U_i}{\partial x_j} + \frac{\partial U_j}{\partial x_i} \right), where \mu = \rho \nu is the dynamic viscosity, and the deviatoric part relates to the strain rate tensor S_{ij} = \frac{1}{2} \left( \frac{\partial U_i}{\partial x_j} + \frac{\partial U_j}{\partial x_i} \right). Under the Boussinesq hypothesis, the turbulent contribution replaces \mu with \mu_t = \rho \nu_t, effectively enhancing the total viscous-like stress while the trace term -\frac{2}{3} k \delta_{ij} ensures the incompressibility of the deviatoric Reynolds stress tensor. This hypothesis simplifies the closure problem in RANS equations by reducing the need to model all six independent Reynolds stress components to determining a single scalar \nu_t. The hypothesis relies on an isotropy assumption for the turbulent viscosity, implying that \nu_t is independent of direction and primarily responds to the mean strain rate. This works well for shear-dominated flows, such as boundary layers, where turbulence is driven mainly by velocity gradients in one direction. However, it fails in flows with strong swirl or separation, where turbulence anisotropy is pronounced, and the linear relation between stresses and strains breaks down due to non-local effects or rapid changes in mean flow topology. The total stress tensor in turbulent flows then combines molecular and turbulent contributions: \tau_{ij}^{\text{total}} = (\nu + \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}, or equivalently, \tau_{ij}^{\text{total}} = \rho (\nu + \nu_t) S_{ij} - \frac{2}{3} \rho k \delta_{ij}, where the term is handled separately in the RANS equations. This form underscores the dominance of eddy viscosity in high-Re regimes, enabling practical simulations while highlighting the need for advanced models in anisotropic conditions.

Prandtl's Mixing-Length Theory

Prandtl's mixing-length theory, proposed by in 1925, offers a foundational semi-empirical framework for estimating the turbulent eddy in shear-dominated flows by analogizing turbulent transfer to in kinetic theory. The core idea posits that turbulent eddies transport fluid parcels over a characteristic distance, termed the mixing length l_m, before the parcels lose their original through mixing with adjacent fluid. This distance l_m serves as a local scale for turbulence, enabling the closure of the Reynolds-averaged Navier-Stokes equations via an effective . In parallel shear flows, the theory expresses the eddy viscosity as \nu_t = l_m^2 \left| \frac{d \bar{U}}{dy} \right|, where \bar{U} is the mean streamwise velocity and y is the coordinate normal to the shear direction; this formulation arises from assuming the fluctuating velocity components scale with the local mean shear times the mixing length, u' \sim l_m \left| \frac{d \bar{U}}{dy} \right|. For wall-bounded flows, Prandtl specified l_m = \kappa y in the near-wall region, with \kappa \approx 0.41 as the von Kármán constant derived from empirical velocity profiles. To extend applicability across the entire boundary layer of thickness \delta, the mixing length is often prescribed as l_m = \kappa y \left(1 - \frac{y}{\delta}\right), which reduces turbulence influence near the outer edge; additionally, the van Driest damping modification, l_m = \kappa y \left[1 - \exp\left(-\frac{y^+}{A^+}\right)\right] with A^+ \approx 26, suppresses excessive mixing in the viscous sublayer close to the wall. This approach operationalizes the Boussinesq eddy viscosity hypothesis specifically for wall-bounded shear layers. It has been applied successfully to flat-plate boundary layers and fully developed pipe flows, where integration of the resulting momentum equation in the logarithmic layer yields the universal velocity profile U^+ = \frac{1}{\kappa} \ln y^+ + B with B \approx 5.1, matching experimental data for high-Reynolds-number turbulent flows. Despite its simplicity and historical impact, Prandtl's theory relies on the assumption of local , where turbulence balances instantaneously, and presumes isotropic mixing perpendicular to the mean . These simplifications render it inadequate for free flows, such as jets or wakes, where the mixing length must instead scale with the flow width rather than wall distance, or for regions with strong adverse pressure gradients that disrupt .

Algebraic Zero-Equation Models

Algebraic zero-equation models, also known as algebraic models, provide a simple approach to closing the Reynolds-Averaged Navier-Stokes (RANS) equations by directly computing the eddy viscosity from local flow variables without solving additional transport equations. These models build on Prandtl's mixing-length theory by specifying algebraic expressions for the mixing length or eddy viscosity, making them particularly suitable for boundary layer flows where the turbulence structure can be prescribed based on assumptions. The Cebeci-Smith model divides the into inner and outer regions to determine the turbulent eddy viscosity \nu_t. In the inner layer, near the wall, \nu_{t,\text{inner}} = l_m^2 |S|, where l_m is the mixing length, typically \kappa y D with van Driest damping D = 1 - \exp(-y^+/26), and |S| is the magnitude of the mean . In the outer layer, \nu_{t,\text{outer}} = 0.0168 U_e \delta^* f(y/\delta), where U_e is the edge velocity, \delta^* is the thickness, \delta is the , and f(y/\delta) is a blending function that decays away from the wall. The total eddy viscosity is taken as the minimum of the inner and outer values to ensure a smooth transition. This formulation was developed for compressible turbulent boundary layers and has been widely applied in computations. The Baldwin-Lomax model (1978) also employs a two-layer structure but incorporates modifications for better handling of wakes and separated flows. For the inner layer, the mixing length is l_m = \kappa y (1 - y/\delta)^{0.5} D, with eddy viscosity computed as \nu_{t,\text{inner}} = l_m^2 |\omega|, where \omega is the magnitude and D is the van Driest damping factor. In the outer layer and wakes, the model shifts to a -based approach, defining \nu_{t,\text{outer}} through a maximum function of times distance from the wall, scaled by empirical constants such as 0.0168 for the . This allows the model to avoid explicit edge detection by relying on local maxima in the profile. These models offer significant advantages in computational efficiency, as they eliminate the need to solve partial differential equations for turbulence quantities, enabling rapid solutions in integral boundary layer methods and early Navier-Stokes solvers. They are particularly effective for attached, equilibrium boundary layers in high-speed flows, such as those in aerodynamics. However, their algebraic nature leads to shortcomings, including a lack of universality across flow regimes; they perform poorly in separated flows, strong adverse pressure gradients, or non-equilibrium conditions like shock-boundary layer interactions, where the assumed local equilibrium fails.

One-Equation Turbulence Models

Spalart-Allmaras Model

The Spalart-Allmaras model is a one-equation turbulence model designed specifically for aerodynamic applications, solving a single for a modified turbulent \tilde{\nu} that relates to the actual turbulent eddy \nu_t through \nu_t = \tilde{\nu} f_{v1}, where f_{v1} = \frac{\chi^3}{\chi^3 + c_{v1}^3} and \chi = \tilde{\nu}/\nu is a blending function ensuring near-wall damping. This approach avoids the numerical and physical challenges associated with solving a transport equation for turbulent kinetic energy in other one-equation models, while employing the Boussinesq eddy hypothesis to model Reynolds stresses. Developed in 1992 by Philippe R. Spalart and Stephen R. Allmaras at Boeing, the model targets attached wall-bounded flows and those with mild separation, such as those encountered in external aerodynamics. The equation for \tilde{\nu} in incompressible form is: \frac{\partial \tilde{\nu}}{\partial t} + \bar{u}_j \frac{\partial \tilde{\nu}}{\partial x_j} = c_{b1} \tilde{S} \tilde{\nu} - c_{w1} f_w \left( \frac{\tilde{\nu}}{d} \right)^2 + \frac{1}{\sigma} \left[ \frac{\partial}{\partial x_j} \left( (\nu + \tilde{\nu}) \frac{\partial \tilde{\nu}}{\partial x_j} \right) + c_{b2} \frac{\partial \tilde{\nu}}{\partial x_i} \frac{\partial \tilde{\nu}}{\partial x_i} \right], where \tilde{S} is a modified incorporating the magnitude of the and a rotation correction term, d is the wall distance, and the production and destruction terms balance to yield the correct asymptotic behavior near walls and in the outer layer. The model includes damping functions like f_w to control the destruction term, preventing excessive turbulence in stagnation regions, and the diffusion term uses both molecular and turbulent viscosity for stability. These features make the robust and relatively simple to implement in structured solvers. Calibration of the model involved adjusting empirical constants—such as c_{b1} = 0.1355, \sigma = 2/3, c_{b2} = 0.622, \kappa = 0.41, c_{v1} = 7.1, c_{w2} = 0.3, and c_{w3} = 2—against experimental data from fully developed flows at moderate Reynolds numbers and equilibrium flat-plate layers, ensuring accurate prediction of skin friction and profiles without sensitivity to values. This tuning prioritized high-Reynolds-number attached flows, resulting in good performance for external aerodynamic simulations like and flows, where it has been widely adopted due to its and low computational cost. However, the model exhibits limitations in free flows, such as jets and wakes, where it overpredicts spreading rates, as the calibration does not fully capture unsteadiness away from walls. The Baldwin-Barth model is a one-equation turbulence model designed for high-Reynolds-number wall-bounded flows, solving a transport equation for the turbulent Reynolds number \tilde{R}_T = k^2 / (\nu \epsilon), where k is the turbulent kinetic energy, \nu is the molecular viscosity, and \epsilon is the dissipation rate. This formulation derives from a simplified version of the standard k-\epsilon model, incorporating an algebraic relation for \epsilon = c k^{3/2}/l with the mixing length l determined near the wall via van Driest damping and in the outer layer based on equilibrium assumptions. The eddy viscosity is computed as \nu_t = c_\mu k^2 / \epsilon with c_\mu = 0.09, ensuring consistency with the Boussinesq hypothesis while avoiding explicit algebraic length-scale specifications in the transport equation. A key feature of the Baldwin-Barth model is its inclusion of history effects through the transport of \tilde{R}_T, which captures non-equilibrium influenced by upstream conditions, such as in separated flows over airfoils. Realizability constraints are enforced on by limiting anti-diffusive numerical fluxes to prevent negative values, particularly requiring adequate grid resolution near where y^+ \approx 1. The model employs functions D_1 = 1 - \exp(-y^+/26) and D_2 = 1 - \exp(-y^+/10) in the eddy viscosity to blend inner and outer layer behaviors smoothly. Related one-equation models include variants of the Spalart-Allmaras approach, such as the SA-neg , which modifies the original transport equation for the modified eddy \tilde{\nu} to handle negative values robustly by switching to a positive-part when \tilde{\nu} < 0, improving numerical stability without altering physics in equilibrium regions. Another example is the Rodi k-l model, often used in near-wall two-layer approaches, which solves a transport equation for k combined with an algebraic length scale l calibrated from direct numerical simulation data to represent viscosity-affected regions up to where \nu_t / \nu \approx 10. These k-based one-equation models offer advantages over zero-equation algebraic baselines by transporting a turbulence scale to account for non-local effects, though they remain simpler than two-equation models. Limitations include sensitivity to inlet turbulence specifications, particularly free-stream \tilde{R}_T values exceeding 100, and reduced accuracy in predicting separation under strong adverse pressure gradients compared to models with full dissipation transport. For instance, computations of transonic airfoil flows show improved lift predictions over algebraic models but discrepancies in pressure distributions attributable to experimental interferences.

Two-Equation Turbulence Models

k-ε Model and Variants

The k-ε model, developed by Launder and Spalding in 1974, is a seminal two-equation turbulence closure approach that solves transport equations for the turbulent kinetic energy k and its dissipation rate \varepsilon, providing a framework to estimate the turbulence length and time scales in Reynolds-Averaged Navier-Stokes (RANS) simulations. It assumes a Boussinesq approximation for the Reynolds stresses, linking them to the mean velocity gradients via an eddy viscosity. This model has become a cornerstone in computational fluid dynamics due to its balance of accuracy and computational cost for engineering applications involving fully turbulent flows away from boundaries. The transport equation for k is given by \frac{Dk}{Dt} = P_k - \varepsilon + \nabla \cdot \left[ \left( \nu + \frac{\nu_t}{\sigma_k} \right) \nabla k \right], where P_k = -\overline{u_i' u_j'} \frac{\partial \overline{U}_i}{\partial x_j} represents the production of turbulent kinetic energy, \nu is the molecular viscosity, and \nu_t is the turbulent viscosity. The corresponding equation for \varepsilon is \frac{D\varepsilon}{Dt} = C_{\varepsilon 1} \frac{\varepsilon}{k} P_k - C_{\varepsilon 2} \frac{\varepsilon^2}{k} + \nabla \cdot \left[ \left( \nu + \frac{\nu_t}{\sigma_\varepsilon} \right) \nabla \varepsilon \right]. The eddy viscosity is modeled as \nu_t = C_\mu \frac{k^2}{\varepsilon}, closing the system under the eddy viscosity hypothesis. These equations incorporate modeled diffusion terms to account for turbulent transport, with production limited to prevent unphysical buildup in low-shear regions. Model constants are calibrated to match experimental data from homogeneous shear flows and grid-generated turbulence decay, ensuring consistency with observed decay rates and shear-induced amplification. Standard values include C_\mu = 0.09, C_{\varepsilon 1} = 1.44, C_{\varepsilon 2} = 1.92, \sigma_k = 1.0, and \sigma_\varepsilon = 1.3, which provide a good fit for a broad class of internal and external flows without excessive tuning. These constants reflect empirical adjustments to the modeled terms, particularly the destruction in the \varepsilon equation, to align with measured turbulence spectra in canonical experiments. Variants of the k-ε model address limitations in specific flow regimes by modifying the transport equations or closures. The RNG k-ε model, derived using renormalization group theory, introduces additional terms to the \varepsilon equation that account for smaller-scale effects and rapid strain, improving predictions in swirling and curved flows like cyclone separators. It analytically derives constants through a perturbative expansion of the Navier-Stokes equations, yielding \alpha = 1.39 for the inverse turbulent Prandtl number and extra dissipation for swirl, enhancing accuracy over the standard model by about 10-20% in rotating shear layers without ad hoc adjustments. The realizable k-ε model, introduced by Shih et al. in 1995, ensures mathematical consistency by enforcing realizability constraints on the Reynolds stresses, such as positive definiteness of the stress tensor. It replaces the constant C_\mu with a variable form dependent on strain and rotation rates, \eta = S \frac{k}{\varepsilon}, where S is the mean strain invariant, allowing C_\mu to range from approximately 0.023 to 0.05. This modification better handles flows with strong rotation or separation, reducing overprediction of normal stresses in planar channels by up to 15% compared to the standard version, while retaining the core transport structure. Low-Reynolds-number extensions, such as those incorporating damping functions near walls, further adapt the model for boundary layers by resolving the viscous sublayer. The standard k-ε model excels in robustness and efficiency for free shear flows, such as round jets and plane mixing layers, where it accurately captures spreading rates within 5-10% of experimental data due to its sensitivity to production-dissipation balance. It is particularly reliable for impinging jets in engineering contexts like combustion chambers, providing stable convergence even on coarse meshes. However, it struggles near solid walls, overpredicting turbulence levels in adverse pressure gradients and requiring wall functions or low-Re modifications to avoid errors exceeding 20% in skin friction predictions. These weaknesses stem from the simplistic \varepsilon destruction term, which inadequately models low-Reynolds-number damping.

k-ω Model and SST Formulation

The k-ω turbulence model, developed by Wilcox in 1988, is a two-equation eddy-viscosity approach that solves transport equations for the turbulent kinetic energy k and the specific dissipation rate \omega, where \omega serves as a scale-determining variable representing the dissipation per unit kinetic energy. This formulation addresses limitations in earlier models by providing better resolution near walls and in adverse pressure gradients without requiring low-Reynolds-number corrections. The turbulent viscosity is defined as \nu_t = k / \omega. The transport equation for k in the basic k-ω model (incompressible form) is given by \frac{Dk}{Dt} = P_k - \beta^* k \omega + \nabla \cdot \left[ \left( \nu + \frac{\nu_t}{\sigma_k} \right) \nabla k \right], and for \omega, \frac{D\omega}{Dt} = \alpha \frac{\omega}{k} P_k - \beta \omega^2 + \nabla \cdot \left[ \left( \nu + \frac{\nu_t}{\sigma_\omega} \right) \nabla \omega \right], where P_k denotes the production of turbulent kinetic energy, \nu is the molecular viscosity, and the model constants are \alpha = 5/9 \approx 0.556, \beta = 3/40 = 0.075, \beta^* = 0.09, \sigma_k = 0.5, and \sigma_\omega = 0.5. These equations stem from a reassessment of the scale-determining equation, emphasizing improved behavior for boundary layers. The variable \omega relates to the dissipation rate \epsilon from \epsilon-based models via \omega = \epsilon / (\beta^* k). Wilcox extended the model to compressible flows, incorporating density variations and heat transfer effects for high-speed applications. The Shear Stress Transport (SST) formulation, developed by Menter in 1994, enhances the k-ω model by blending it with a k-ε model to improve performance in both near-wall and free-shear regions. This blending uses a switching function F_1 = \tanh(\arg_1^4), where \arg_1 depends on the distance to the nearest wall, ensuring the k-ω behavior dominates in the inner layer (F_1 \approx 1) and a transformed k-ε formulation activates in the outer region (F_1 \approx 0). Model constants and terms are interpolated as \phi = F_1 \phi_1 + (1 - F_1) \phi_2, where subscripts 1 and 2 denote k-ω and k-ε values, respectively. A key feature of the SST model is a limiter on the eddy viscosity to better capture the transport of principal shear stress, formulated as \nu_t = \frac{k}{\max(a_1 \omega, |S| F_2)}, where |S| is the invariant measure of the strain rate, a_1 = 0.31, and F_2 is another blending function; this effectively limits C_\mu dynamically for improved separation prediction. The SST approach enables automatic wall treatment, as it integrates low-Reynolds-number effects near boundaries without damping functions. It excels in simulating adverse pressure gradients and transitional flows, making it particularly suitable for turbomachinery applications.

Performance Comparisons

Performance comparisons of two-equation turbulence models, particularly the k-ε and k-ω/SST formulations, are often evaluated using standard benchmarks such as fully developed channel flow, backward-facing step flows, and airfoil stall scenarios. In channel flow, both models generally reproduce the logarithmic law of the wall with reasonable accuracy for high-Reynolds-number conditions, though the k-ε model aligns closely with experimental velocity profiles in the log-law region due to its robust treatment of free-stream turbulence. The k-ω/SST model performs comparably but may require finer near-wall resolution to avoid sensitivity to freestream values. For the backward-facing step benchmark, which tests separation and reattachment, the standard k-ε model tends to underpredict the reattachment length (typically by 20-30% compared to experiments like Driver and Seegmiller's), leading to inaccurate separation bubble sizes, while the k-ω/SST model provides better agreement with measured reattachment points due to its improved near-wall behavior. In airfoil stall prediction, such as for the NACA0012, the k-ε model often overestimates lift in post-stall regimes with errors up to 15-20% in drag coefficients, whereas k-ω/SST captures the onset of separation more accurately, reducing lift prediction errors by approximately 10-15% through better handling of adverse pressure gradients. The k-ε model excels in internal flows like diffusers and pipes, where it reliably predicts bulk flow features without excessive sensitivity to inlet conditions, but it struggles with separation-dominated cases by overpredicting shear stresses in recirculating regions. In contrast, the k-ω/SST formulation offers superior reattachment predictions and 15-25% improvements in skin friction coefficient accuracy for separated flows, attributed to its blending of k-ω near walls and k-ε in the freestream. Computationally, both models impose similar overheads, roughly 5-15 times that of laminar simulations on equivalent meshes, primarily due to the additional transport equations and near-wall refinements required for turbulence resolution. The k-ε model is generally less demanding than k-ω/SST, especially with wall functions, but both exhibit sensitivity to mesh quality: low-Reynolds-number formulations demand y+ < 1 for accurate viscous sublayer resolution, while wall-function approaches tolerate y+ ≈ 30-100, though exceeding this can degrade log-law compliance. Recent assessments through 2025 highlight the SST model's strengths in hypersonic applications, where modifications like compressibility corrections enhance heat flux and separation predictions in shock-boundary layer interactions by up to 20% over baseline k-ε. Emerging machine learning hybrids are also showing promise for calibrating SST coefficients, improving benchmark accuracies in transitional separations without altering core equations.

Reynolds Stress Models

Transport Equations for Reynolds Stresses

In Reynolds stress models, the transport equations for the individual components of the Reynolds stress tensor R_{ij} = \overline{u_i' u_j'}, which arises from the Reynolds decomposition in the derivation of the Reynolds-averaged Navier-Stokes (RANS) equations, are solved to directly account for the anisotropy of turbulent fluctuations without relying on the Boussinesq approximation of isotropic eddy viscosity. These equations are obtained by taking the appropriate moments of the instantaneous Navier-Stokes equations and averaging, resulting in an exact but unclosed form that requires modeling for practical implementation. The exact transport equation for R_{ij} in incompressible flow, neglecting viscous diffusion at high Reynolds numbers, is given by \frac{DR_{ij}}{Dt} = P_{ij} + \Phi_{ij} + D_{ij} - \varepsilon_{ij} + \pi_{ij}, where \frac{D}{Dt} = \frac{\partial}{\partial t} + \bar{U}_k \frac{\partial}{\partial x_k} is the substantial derivative following the mean flow. The production term P_{ij} represents the transfer of mechanical energy from the mean flow to the turbulent fluctuations and is exactly closed as P_{ij} = -R_{ik} \frac{\partial \bar{U}_j}{\partial x_k} - R_{jk} \frac{\partial \bar{U}_i}{\partial x_k}, where \bar{U}_i is the mean velocity; this term is always positive semi-definite and drives the growth of turbulence in shear flows. The dissipation tensor \varepsilon_{ij} accounts for the irreversible conversion of turbulent kinetic energy to internal energy via viscous effects and is modeled in high-Reynolds-number flows as isotropic, \varepsilon_{ij} = \frac{2}{3} \varepsilon \delta_{ij}, where \varepsilon = \frac{1}{2} \varepsilon_{kk} is the scalar dissipation rate, often obtained from a separate modeled transport equation. The remaining terms require closure approximations. The pressure-strain correlation \Phi_{ij}, which redistributes energy among the Reynolds stress components without altering the trace (turbulent kinetic energy k = \frac{1}{2} R_{ii}), is modeled using proposals from , such as the form \Phi_{ij} = \Phi_{ij}^{(1)} + \Phi_{ij}^{(2)}, where the slow part \Phi_{ij}^{(1)} depends on the anisotropy tensor and the rapid part \Phi_{ij}^{(2)} responds to mean strain; typical coefficients include C_1 = 1.8 and C_2 = 0.6 for the isotropization of production variant. The turbulent diffusion term D_{ij} and pressure-diffusion term \pi_{ij} together represent the spatial transport of R_{ij} by turbulent fluctuations and pressure fluctuations, respectively, and are commonly modeled via the generalized gradient diffusion hypothesis as D_{ij} + \pi_{ij} with C_s \approx 0.22. The system comprises seven independent transport equations for the symmetric R_{ij} components, as the trace equation recovers the modeled transport equation for k; this direct resolution of the full stress tensor enables RSM to capture complex flow features like secondary strains and swirl, where simpler models assuming R_{ij} = -\frac{2}{3} k \delta_{ij} + 2k \nu_t S_{ij} (with S_{ij} the mean strain rate) fail.

Realizability and Nonlinear Extensions

Realizability constraints in Reynolds stress models (RSMs) ensure that the modeled Reynolds stress tensor \overline{u_i u_j} remains physically admissible, meaning it must be symmetric, positive semi-definite (with non-negative eigenvalues), and satisfy the Cauchy-Schwarz inequality |\overline{u_i u_j}| \leq \sqrt{\overline{u_i^2} \overline{u_j^2}} for all components. These conditions prevent unphysical behaviors such as negative turbulent kinetic energy or excessive off-diagonal stresses that could arise from closure approximations. Violations often occur in regions of strong strain or rapid distortion, where linear models fail to maintain tensor integrity; thus, realizability is typically enforced by bounding the pressure-strain correlation term \Phi_{ij} and dissipation tensor through inequality-constrained formulations. Speziale (1991) demonstrated theoretically that many early RSM closures, including those based on the Launder-Reece-Rodi model, can lead to realizability violations in homogeneous shear flows by producing negative normal stresses. To address this, subsequent models incorporate explicit constraints, such as tensorial Schwarz inequalities, directly into the evolution equations for \overline{u_i u_j}, ensuring long-time asymptotic realizability without ad hoc clipping. These enhancements improve stability in simulations of transitional or highly anisotropic turbulence while preserving the model's ability to capture history effects. Nonlinear extensions of RSMs advance beyond the Boussinesq eddy-viscosity hypothesis by modeling the pressure-strain correlation \Phi_{ij} as a nonlinear function of the mean strain-rate tensor S_{kl} and rotation-rate tensor \Omega_{kl}, enabling the prediction of non-equilibrium states where stresses do not align with the principal strain axes. The Speziale-Sarkar-Gatski (SSG) model (1991) exemplifies this approach, employing an invariant dynamical systems framework to construct \Phi_{ij} with quadratic terms in S_{kl} and \Omega_{kl} for the rapid part, and linear-qu quadratic for the slow part, ensuring both realizability and frame-indifference. This formulation, calibrated against homogeneous turbulence data, enhances accuracy in curved and swirling flows by accounting for pressure diffusion and redistribution effects more faithfully than linear models. Low-Reynolds-number extensions integrate RSMs into the near-wall region by introducing damping functions that modulate production, dissipation, and diffusion terms to capture viscous damping without logarithmic wall functions, allowing resolution down to the viscous sublayer. These functions, often derived from direct numerical simulation data, suppress turbulence intensities asymptotically as y^+ \to 0, where y^+ is the wall-normal distance in viscous units. A prominent example is the SSG/LRR-\omega hybrid, which blends the SSG pressure-strain model in the outer flow with the Launder-Reece-Rodi (LRR) variant near walls for improved anisotropy prediction, while replacing the \varepsilon-equation with a transport equation for specific dissipation \omega to reduce stiffness and enhance near-wall robustness. This setup, developed within the European FLOMANIA project (Eisfeld, 2006) and refined in subsequent calibrations, performs well in internal flows with adverse pressure gradients. Despite these improvements, RSMs present significant challenges, including the calibration of 12 to 14 empirical constants per model variant, which demands extensive homogeneous and inhomogeneous flow data for tuning. Numerical stiffness arises from the coupled transport equations for the six independent stress components plus the scale equation, often requiring implicit solvers and fine near-wall grids, increasing computational cost by factors of 10 to 20 compared to k-\varepsilon models. While excelling in flows with strong streamline curvature, rotation, or secondary motions—such as cyclone separators or curved ducts—RSMs remain less adopted in industrial CFD due to these demands, though they offer superior fidelity over eddy-viscosity models in non-equilibrium scenarios.

Large Eddy Simulation Techniques

Spatial Filtering and Subgrid Stresses

In large eddy simulation (LES), spatial filtering is applied to the Navier-Stokes equations to separate the turbulent flow field into large-scale, resolvable components and small-scale, subgrid-scale (SGS) components that must be modeled. The filtering operation, denoted as \tilde{u} = G * u, where G is the filter kernel and * represents convolution, produces the filtered velocity field \tilde{u}, which captures the energy-containing eddies while smoothing out smaller scales. Filters can be implemented explicitly, by directly convolving the velocity field with a specified kernel such as the box filter (top-hat) or Gaussian filter, or implicitly, where the numerical grid acts as the low-pass filter without additional convolution steps. The box filter, defined as G(x) = 1/\Delta for |x| < \Delta/2 and zero otherwise, averages over a finite volume, while the Gaussian filter G(x) = (6/(\pi \Delta^2))^{1/2} \exp(-6 x^2 / \Delta^2) provides smoother decay in Fourier space, reducing Gibbs oscillations but requiring more computational effort. Explicit filtering allows precise control over the cutoff scale but increases costs, whereas implicit filtering leverages the discretization scheme for efficiency, though it may introduce numerical dissipation. The subgrid-scale stress tensor arises from the nonlinear convection term after filtering and is defined as \tau_{ij} = \overline{u_i u_j} - \tilde{u}_i \tilde{u}_j, where the overbar denotes the filtering operation, representing the effect of unresolved scales on the resolved flow. This tensor must be modeled to close the equations, as it encapsulates the transfer of energy and momentum across the filter scale. To analyze its structure, the SGS stress can be decomposed using Leonard's formulation into three components: the Leonard stress L_{ij} = \widetilde{\tilde{u}_i \tilde{u}_j} - \tilde{u}_i \tilde{u}_j, the cross stress C_{ij} = \widetilde{\tilde{u}_i u_j'} + \widetilde{u_i' \tilde{u}_j}, and the Reynolds SGS stress R_{ij} = \overline{u_i' u_j'} - \tilde{u}_i' \tilde{u}_j', where primes denote deviations from the filtered field; this decomposition highlights interactions between resolved and unresolved motions. The Germano identity provides a framework for dynamic modeling procedures by relating subgrid stresses at different filter levels: for a test filter \hat{G} wider than the grid filter G, it states T_{ij} - \hat{\tau}_{ij} = \widehat{\tilde{u}_i \tilde{u}_j} - \hat{\tilde{u}}_i \hat{\tilde{u}}_j = L_{ij}, where T_{ij} is the test-level SGS stress, enabling least-squares minimization to determine model coefficients without a priori assumptions. Applying the filter to the incompressible Navier-Stokes equations yields the filtered momentum equation: \frac{\partial \tilde{u}_i}{\partial t} + \frac{\partial (\tilde{u}_i \tilde{u}_j)}{\partial x_j} = -\frac{1}{\rho} \frac{\partial \tilde{p}}{\partial x_i} + \nu \frac{\partial^2 \tilde{u}_i}{\partial x_j \partial x_j} - \frac{\partial \tau_{ij}}{\partial x_j}, where the SGS stress divergence acts as an additional forcing term that models the influence of small scales on large ones. The filter width \Delta is chosen to align with the integral scale of large eddies, much larger than the \eta where \Delta \gg \eta, ensuring that the resolved scales capture the anisotropic, energy-producing motions while isotropic small scales are modeled. This scale separation enables LES to simulate flows at Reynolds numbers up to approximately $10^5, with computational cost scaling as O(Re) due to a grid resolution proportional to the large-eddy scale, compared to direct numerical simulation (DNS) requiring O(Re^{9/4}) operations to resolve all scales down to \eta.

Smagorinsky Subgrid Model

The provides an eddy viscosity closure for the subgrid-scale (SGS) stresses in (LES), representing the effect of unresolved turbulent scales on the filtered velocity field. Introduced by in his pioneering work on , the model assumes that the SGS turbulence can be characterized by a local mixing length proportional to the computational grid size.091<0099:GCEWTP>2.3.CO;2) The core of the model is the expression for the anisotropic part of the SGS stress tensor: \tau_{ij} - \frac{1}{3} \tau_{kk} \delta_{ij} = -2 (C_s \Delta)^2 |\tilde{S}| \tilde{S}_{ij}, where \tilde{S}_{ij} is the filtered , |\tilde{S}| = \sqrt{2 \tilde{S}_{mn} \tilde{S}_{mn}} is its magnitude, C_s \approx 0.18 is an empirical constant calibrated for isotropic , and \Delta = (\Delta_x \Delta_y \Delta_z)^{1/3} defines the filter width from the local spacing. This formulation implies an SGS eddy of \nu_{sgs} = (C_s \Delta)^2 |\tilde{S}|, which scales the transfer of energy from resolved to subgrid scales based on the local deformation of the large eddies.091<0099:GCEWTP>2.3.CO;2) The model derives from Prandtl's mixing-length analogy, setting the characteristic length scale of SGS eddies as l = C_s \Delta, such that the turbulent viscosity follows \nu_t = l^2 |\tilde{S}|. It relies on the assumption of local equilibrium between SGS kinetic energy production P_{sgs} and dissipation \epsilon_{sgs}, implying that the energy cascade rate is balanced at subgrid scales without net accumulation. This equilibrium hypothesis simplifies the closure but aligns with Kolmogorov's theory for the inertial subrange in homogeneous turbulence.091<0099:GCEWTP>2.3.CO;2) While computationally efficient and robust for fully developed turbulent flows, the Smagorinsky model exhibits limitations, such as overdissipation during laminar-to-turbulent transitions and reduced accuracy in complex flows with or separation. Near walls, it overpredicts due to the fixed length scale, necessitating a damping function like f_w = [1 + 50 (\Delta / d)^2]^{-1.5}, where d is the to , to reduce the effective \nu_{sgs} in the viscous sublayer.

Dynamic and Localized Models

The dynamic Smagorinsky model addresses limitations of the static formulation by computing the Smagorinsky constant C_s adaptively from the resolved flow field, enhancing accuracy in inhomogeneous and transitional . Introduced by Germano et al. in 1991, it employs the Germano , which relates subgrid-scale (SGS) stresses at the grid level \tau_{ij} to those at a coarser test level T_{ij}. Lilly refined this in 1992 using a least-squares minimization to determine the model coefficient, ensuring stability and robustness. The core of the model involves the resolved stress tensor L_{ij} = \widehat{\bar{u}_i \bar{u}_j} - \hat{\bar{u}}_i \hat{\bar{u}}_j, where \bar{} denotes the grid filter and \hat{} the test filter, typically with scale ratio \gamma = 2. This L_{ij} is expressed as L_{ij} = T_{ij} - \hat{\tau}_{ij} = C M_{ij}, with M_{ij} constructed from the difference in Smagorinsky eddy viscosities between scales. The coefficient C = \langle L_{ij} M_{ij} \rangle / \langle M_{ij} M_{ij} \rangle is evaluated locally and then averaged over homogeneous directions to mitigate noise from insufficient sampling, often with clipping $0 \leq C \leq C_{\max} for . Scale-dependent dynamic procedures extend this framework by allowing C_s to vary across scales, improving performance in flows with broad inertial ranges, such as atmospheric boundary layers. Localized models avoid global averaging, enabling pointwise computation suitable for complex geometries and non-homogeneous . The scale-similarity model, proposed by Bardina et al. in 1980, approximates SGS stresses via similarity to resolved scales: \tau_{ij} \approx \bar{u}_i \bar{u}_j - \tilde{u}_i \tilde{u}_j, where \tilde{} is a secondary , capturing and scale interactions without eddy . Nonlinear extensions, such as those incorporating higher-order tensor representations, enhance capture in sheared flows. The Vreman model (2004) provides a localized eddy- alternative, defining \nu_t = 2.5 B_\beta / |\mathbf{S}|^2 with B_\beta = B_{\beta 11} + B_{\beta 22} + B_{\beta 33}, where terms like B_{\beta 11} = \Delta^2 (\partial_k \bar{u}_i \partial_k \bar{u}_i)(\partial_j \bar{u}_i \partial_j \bar{u}_i) - (\partial_k \bar{u}_i \partial_j \bar{u}_i)^2 detect local and vanish in laminar regions, outperforming Smagorinsky in transitional shear flows. Post-2000 advances include implicit large-eddy simulation (ILES), where numerical dissipation from high-order schemes substitutes for explicit SGS modeling, proving effective for hypersonic boundary layers by implicitly handling shocks and turbulence without closures.

Hybrid and Emerging Methods

Detached Eddy Simulation

(DES) is a hybrid turbulence modeling approach that combines Reynolds-Averaged Navier-Stokes (RANS) methods in near-wall regions with (LES) in separated flow areas, enabling efficient simulations of high-Reynolds-number flows with massive separation. Proposed by Spalart et al. in 1997, DES modifies the Spalart-Allmaras one-equation RANS model by altering the dissipation term to promote a switch to LES mode away from walls. Specifically, the model employs RANS modeling near solid surfaces, transitioning to LES in detached layers where large-scale eddies dominate and grid spacing supports . The key modification involves replacing the standard dissipation \tilde{\epsilon} = \rho \tilde{\nu} / l_{RANS}^2 with \tilde{\epsilon} = \rho k^{3/2} / l_{DES}, where the DES length scale is defined as l_{DES} = \min(l_{RANS}, C_{DES} \Delta), with C_{DES} \approx 0.65 and \Delta as the local spacing. This formulation ensures that in attached boundary layers, l_{RANS} < C_{DES} \Delta, maintaining RANS behavior, while in separated regions with finer grids, the LES mode activates by limiting the eddy based on grid . Subsequent developments addressed early limitations of DES, particularly its sensitivity to grid resolution in boundary layers, which could prematurely trigger LES mode and cause premature separation. In 2006, Spalart et al. introduced Delayed Detached Eddy Simulation (DDES) by replacing the wall distance d in the length scale with a blended distance d_b = d - f_d \max(d, C_{DES} \Delta), where f_d is a blending function that approaches zero near walls and one in the logarithmic layer. This modification prevents grid-induced separation (GIS) by shielding boundary layers from the DES limiter, allowing RANS to persist even on finer near-wall grids. Further refinement came with Improved Delayed Detached Eddy Simulation (IDDES), which enhances blending through additional functions to better handle both attached and separated flows, incorporating protections against modeled stress depletion and improving transitions in shear layers. DES variants have found widespread application in , particularly for simulating aircraft wakes, landing gear noise, and massively separated flows around wings at high angles of attack. For instance, has been used to predict unsteady in configurations, capturing and wake instabilities with accuracy superior to steady RANS while requiring computational costs intermediate between RANS and full —typically 10 to 100 times higher than RANS but far less than wall-resolved . These methods excel in scenarios involving complex geometries and transient phenomena, such as bluff-body wakes or interactions. Despite these advances, DES approaches retain limitations, including the modeled stress depletion (MSD) issue, where excessive grid refinement near walls reduces modeled Reynolds stresses without sufficient resolved turbulent to compensate, leading to underprediction of skin friction and delayed separation. DDES and IDDES mitigate MSD through shielding functions, but the methods still demand fine, isotropic grids in LES regions—often with \Delta \approx 0.1 to $1\% of the boundary layer thickness—to resolve large eddies accurately, increasing computational expense for industrial-scale simulations.

Machine Learning Applications in Modeling

Machine learning has emerged as a powerful tool to address the closure problem in turbulence modeling by leveraging high-fidelity , such as from direct numerical simulations (DNS), to learn complex relationships that traditional physics-based models approximate. Neural serve as surrogates for closure coefficients, particularly in Reynolds-averaged Navier-Stokes (RANS) models, where convolutional neural (CNNs) are trained on DNS to predict eddy viscosity \nu_t or Reynolds stress anisotropies directly from resolved features. For instance, structured neural have demonstrated improved accuracy in predicting turbulent and stresses in channel by eliminating the need for additional equations, achieving correlation coefficients exceeding 0.96 in a priori tests against DNS benchmarks. Physics-informed neural networks (PINNs) enhance these surrogates by incorporating Navier-Stokes constraints and conservation laws into the loss function, ensuring physical consistency and aiding generalization across flow regimes. In turbulent flow simulations, PINNs have been used to calibrate RANS model parameters, such as in the k-\epsilon model, using DNS , improving prediction accuracy. For (LES), generative adversarial networks (GANs) predict subgrid-scale (SGS) stresses by learning mappings from coarse-grained fields to fine-scale dynamics, capturing non-local effects like backscattering more effectively than eddy-viscosity models; physics-informed enhanced super-resolution GANs, for example, have shown superior reconstruction of turbulent channel flow structures with mean absolute errors below 5% in profiles. (RL) further advances wall modeling in LES, where multi-agent RL frameworks develop data-driven wall models for turbulent boundary layers, achieving low errors in velocity across high Reynolds numbers. Recent developments from 2020 to 2025 have integrated these techniques into practical frameworks, notably through NASA's (TMR), where augments RANS models for transition prediction using field inversion and CNN-based surrogates on cases like the HIFiRE-1 experiment. In hypersonic applications, ML-calibrated Reynolds stress models (RSMs) have reduced errors in boundary layer transition predictions by approximately 40% for Mach 4.5 flows over swept wings, by learning N-factor envelopes from correlations derived from DNS. These advancements, exemplified in seminal works like Kochkov et al.'s universal ML approximations for Navier-Stokes solutions, underscore ML's potential for high-impact contributions in simulations. Despite these gains, challenges persist in ML applications to turbulence modeling. Generalization beyond training flows remains limited, as models trained on specific DNS datasets often underperform in unseen regimes due to the non-universal nature of turbulence statistics, with extrapolation errors exceeding 20% in out-of-distribution tests like varying Reynolds numbers. High-fidelity DNS data requirements are prohibitive, necessitating vast computational resources to span the turbulence parameter space adequately. Interpretability issues also hinder adoption, as black-box neural architectures obscure causal physical mechanisms, complicating validation against empirical models and integration into certified engineering workflows.

References

  1. [1]
    Turbulence Modeling Resource
    ### Definition and Overview of Turbulence Modeling (NASA Perspective)
  2. [2]
    Turbulence modeling and simulation advances in CFD during the ...
    This paper is a short retrospective review of the predictive methods of turbulent flows in Computational Fluid Dynamics over the last 50 years
  3. [3]
    Review of Challenges and Opportunities in Turbulence Modeling
    Turbulence modeling is a fluid mechanics method that uses numerical computation and analysis to study complex turbulent flows in fluid systems [1]. There are ...
  4. [4]
  5. [5]
  6. [6]
    Turbulence Modeling; Uncertainty Quantification; Computational ...
    Sep 4, 2025 · Turbulence Models represent the workhorse for simulations used in engineering design and analysis. Despite their low computational cost and ...Missing: scholarly | Show results with:scholarly
  7. [7]
    [PDF] Lecture 2: The Navier-Stokes Equations - Projects at Harvard
    Sep 9, 2015 · Summarizing, the Navier-Stokes equations of continuum fluid mechanics are ”simply” Newton's law ma = F as applied to a small volume of fluid.
  8. [8]
  9. [9]
    [PDF] A Primer on Direct Numerical Simulation of Turbulence - ePrints Soton
    Direct Numerical Simulation (DNS) is the branch of CFD devoted to high-fidelity solution of turbulent flows. DNS differs from conventional CFD in that the ...
  10. [10]
    [PDF] Chapter 3 A statistical description of turbulence
    A statistical measure is an average of some kind: over the symmetry coordinates, if any are available (e.g, a time average for stationary flows); over multiple ...
  11. [11]
    [PDF] Reynolds Stresses
    These additional terms in the governing equations for turbulent flow are called Reynolds stress terms for the following reason. It is noticeable that the ...
  12. [12]
    On the convergence and scaling of high-order statistical moments in ...
    Dec 21, 2017 · The aim of the current study is to revise statistical moments in turbulent pipe flow, such as turbulent intensities, skewness, and flatness ...
  13. [13]
    [PDF] Chapter 6 Isotropic homogeneous 3D turbulence
    Hence the energy spectrum has the information content of the two-point correlation. E(k) contains directional information. More usually, we want to know the ...
  14. [14]
    [PDF] Homogeneous turbulence: an introductory review - DAMTP
    Sep 6, 2012 · 'space average'; ergodic theory tells us that these two averages are, under the condition of homogeneity, equivalent. Turbulence that is ...
  15. [15]
    IV. On the dynamical theory of incompressible viscous fluids and the ...
    Reynolds Osborne. 1895IV. On the dynamical theory of incompressible viscous fluids and the determination of the criterionPhilosophical Transactions of the ...
  16. [16]
    [PDF] Reynolds Averaging
    Mechanics of turbulence. MIT Press, Cambridge, Mass., 769 pp. Reynolds, Osborne, 1895: On the Dynamical Theory of Incompressible Viscous Fluids and the.
  17. [17]
    The mechanics of an organized wave in turbulent shear flow
    Mar 29, 2006 · Journal of Fluid Mechanics , Volume 41 , Issue 2 , 13 April 1970 ... A. K. M. F. Hussain (a1) and W. C. Reynolds (a1); DOI: https://doi.org ...
  18. [18]
    [PDF] The Turbulence Closure Problem and Derivation of the Reynolds ...
    Mar 20, 2008 · τ ≡ − in the momentum equation represents six additional unknowns! Note that these equations are exact. (within the limits of the ...
  19. [19]
    [PDF] Introduction
    Statistical studies of the equations of motion always lead to a situation in which there are more unknowns than equations. This is called the closure problem of ...
  20. [20]
    [PDF] 6 Turbulence - DAMTP
    We find that we have an infinite hierarchy of equations. This is not unusual ... Rearranging, we have an expression for the three-point correlations,.
  21. [21]
    [PDF] First- and Second-Moment Equations - Colorado State University
    Mellor and Yamada (1974) presented a hierarchy of turbulence closure models, ranging form a fully prognostic system of second-moment equations (Level 4) to a ...
  22. [22]
    [PDF] REPORT ONINVESTIGATION O,FDEVELOPED TURBULENCE By L ...
    turbulent “exchange” and in its Umension, which is the ssme as that of V, it is the ~roduct of a length snd a velocity. The velocity is the transverse velocity ...<|separator|>
  23. [23]
    [PDF] The Local Structure of Turbulence in Incompressible Viscous Fluid ...
    The first hypothesis of similarity. For the locally isotropic turbulence the distributions Fn are uniquely determined by the quantities v and e. The ...
  24. [24]
    The Contributions of A. N. Kolmogorov to the theory of turbulence
    PDF | Two of the papers published by Kolmogorov in 1941 are generally considered to be the origin of modern turbulence theory, including the concepts of.
  25. [25]
  26. [26]
    Welcome to AE Resources
    The Reynolds stresses introduced six new unknowns into the problem. Several methods could be used to estimate this tensor: Adding more equations to the problem ...<|separator|>
  27. [27]
    Spalart-Allmaras Model - Turbulence Modeling Resource
    Linear models use the Boussinesq assumption for the constitutive relation: \tau_{ij} = 2 \mu_t \left(S_{ij}. where the last term is generally ignored for ...
  28. [28]
    [PDF] About Boussinesq's turbulent viscosity hypothesis - HAL
    ... Boussinesq's original publication, which can be found in French at Gallica's web site (http://gallica.bnf.fr). In his 1877 publication, Boussinesq performs ...
  29. [29]
    Boussinesq's Hypothesis - an overview | ScienceDirect Topics
    7.1.5 Eddy-viscosity hypothesis. One of the most significant contributions to turbulence modeling was presented in 1877 by Boussinesq [17, 18] ...Missing: original | Show results with:original<|control11|><|separator|>
  30. [30]
    SIMULATION OF SWIRLING FLOWS BASED ON MODIFIED TWO ...
    Linear RANS models based on the Boussinesq hypothesis are not able to give satisfactory results for such problems, because this hypothesis assumes isotropic ...Missing: limitations | Show results with:limitations<|control11|><|separator|>
  31. [31]
    Turbulence model form errors in separated flows | Phys. Rev. Fluids
    Feb 15, 2023 · In turbulence modeling, model form errors are introduced in the prediction of the Reynolds stresses through the Boussinesq hypothesis and ...
  32. [32]
    [PDF] Turbulence Models and Their Application to Complex Flows R. H. ...
    Turbulence is a three-dimensional unsteady viscous phenomenon that occurs at high Reynolds number. Turbulence is not a fluid property, but is a property of the.
  33. [33]
  34. [34]
    Analysis of turbulent boundary layers : Cebeci, Tuncer
    Oct 6, 2022 · Publication date: 1974 ; Topics: Turbulent boundary layer ; Publisher: New York, Academic Press ; Collection: internetarchivebooks; printdisabled.
  35. [35]
    Thin-layer approximation and algebraic model for separated ...
    Aug 16, 2012 · Extension of a simplified Baldwin–Lomax model to nonequilibrium turbulence: Model, analysis and algorithms ... 16 January 1978 - 18 January 1978.
  36. [36]
    [PDF] A One-Equation Turbulence Model for Aerodynamic Flows
    Manuscript received March 27, 1992; accepted May 3, 1993. Résumé. Abstract. Spalart P. R., Allmaras S. R., La Recherche Aérospatiale, 1994, n° 1, ...
  37. [37]
  38. [38]
    Baldwin-Barth model - CFD Online
    Jun 12, 2007 · Baldwin, B.S. and Barth, T.J. (1990), A One-Equation Turbulence Transport Model for High Reynolds Number Wall-Bounded Flows, NASA TM 102847.
  39. [39]
    [PDF] Modifications and Clarifications for the Implementation of the Spalart ...
    The Spalart-Allmaras (S-A) turbulence model [1, 2] has been widely used and has proven to be numerically well behaved in most cases.<|control11|><|separator|>
  40. [40]
    Experience with two-layer models combining the k ... - AIAA ARC
    Experience with two-layer models combining the k-epsilon model with a one-equation model near the wall. W. RODI. W. RODI.
  41. [41]
    Reassessment of the scale-determining equation for advanced ...
    Reassessment of the scale-determining equation for advanced turbulence models. David C. Wilcox. David C. Wilcox. DCW Industries, Inc., La Canada, California.
  42. [42]
    Wilcox k-omega Model - Turbulence Modeling Resource
    The Wilcox k-omega Turbulence Model. This web page gives detailed information on the equations for various forms of the Wilcox k-omega turbulence model.
  43. [43]
    Two-equation eddy-viscosity turbulence models for engineering ...
    Volume 32, Issue 8. Full Access. Two-equation eddy-viscosity turbulence models for engineering applications. Published Online:17 May 2012.
  44. [44]
    Menter Shear Stress Transport Model
    All forms of the model given on this page are linear eddy viscosity models. Linear models use the Boussinesq assumption for the constitutive relation: \tau_{ij} ...Missing: hypothesis | Show results with:hypothesis
  45. [45]
    Which Turbulence Model Should I Choose for My CFD Application?
    Jul 6, 2017 · The low Reynolds number k-ε model is similar to the k-ε model, but does not need wall functions: it can solve for the flow everywhere. It is a ...
  46. [46]
    K-Omega Turbulence Models | Global Settings - SimScale
    The k − ω SST model is the most widely accepted industry standard turbulence model. It provides a better prediction of flow separation than most RANS models and ...K-Omega Turbulence Models · Transport Equations for the... · Applications
  47. [47]
    Turbulent Flow Past a Backward-Facing Step: A Critical Evaluation ...
    Aug 5, 2025 · It is shown in this paper that the original version of the renormalization group K-epsilon model substantially underpredicts the reattachment ...
  48. [48]
    Assessment of Turbulence Models on a Backward Facing Step Flow ...
    It was found that results from the first-order standard k-epsilon and Realizable k-epsilon model were better than the second-order Reynolds Stress. Transport ...
  49. [49]
    Prediction of Airfoil Stall Based on a Modified k−v2¯−ω Turbulence ...
    Jan 16, 2022 · The SST turbulence model gives a relatively accurate separation point, but it underpredicts the level of turbulence stresses in the separated ...
  50. [50]
    [PDF] Comparative Study on Effective Turbulence Model for NACA0012 ...
    The results obtained have shown that the Standard K – epsilon model is found to have less error percentage in comparison to other turbulence models. The count ...<|control11|><|separator|>
  51. [51]
    (PDF) The Evaluation of k-ε and k-ω Turbulence Models in ...
    Aug 6, 2025 · The main objective is to evaluate the performance of the realizable k-ε and k-ω SST models qualitatively and quantitatively in modelling flow of a highly bend ...
  52. [52]
    [PDF] Development and Application of Elliptic Blending Lag k-omega SST ...
    May 15, 2020 · 3.2.5 Flow past a Curved Backward-Facing Step. The curved backward-facing step is a more complex flow separation case; it has been studied by ...<|separator|>
  53. [53]
    [PDF] Understanding the Role of CFD, Including RANS, DES, LES and ...
    Apr 28, 2022 · – Computational cost is excessive (at least 100 times that of RAS). • Expected trend is to continue leveraging RAS capabilities in design and ...
  54. [54]
    (PDF) COMPARISON OF STANDARD k-epsilon AND SST k-omega ...
    The findings indicate that while both models yield close results, the standard k-ε model requires less computational power than the SST k-ω model. What does the ...
  55. [55]
    What Y+ values should be used for the ke and SST turbulence ...
    Oct 8, 2023 · SST has no wall function and so the Y+ needs to be lower. A value of 2 is typically OK although this might need to be as low as 1, especially ...
  56. [56]
    Enhancing the SST turbulence model for predicting heat flux in ...
    The outcomes demonstrate a substantial improvement in the model's ability to predict heat transfer in hypersonic SWTBLI flows without compromising the baseline ...
  57. [57]
    Machine Learning Based Turbulence Model Calibration for ...
    Jan 3, 2025 · In this work, we aim to develop a deep learning surrogate model to optimize the coefficients of the Shear Stress Transport (SST) turbulence ...
  58. [58]
    Progress in the development of a Reynolds-stress turbulence closure
    Mar 29, 2006 · The paper develops proposals for a model of turbulence in which the Reynolds stresses are determined from the solution of transport equations.
  59. [59]
    On the realizability of reynolds stress turbulence closures
    The realizability of Reynolds stress models in homogeneous turbulence is critically assessed from a theoretical standpoint.
  60. [60]
    SSG/LRR Full Reynolds Stress Model
    This web page gives detailed information on the equations for various versions of the blended SSG/LRR full second-moment Reynolds stress models.
  61. [61]
    [PDF] a review of reynolds stress models for turbulent shear flows
    A detailed review of recent developments in Reynolds stress modeling for incompressible turbulent shear flows is provided. The mathematical foundations.
  62. [62]
    Energy Cascade in Large-Eddy Simulations of Turbulent Fluid Flows
    Journals & Books · View PDF; Download full volume. Search ScienceDirect. Elsevier. Advances in Geophysics · Volume 18, Part A, 1975, Pages 237-248. Advances in ...
  63. [63]
    Turbulence: Subgrid-Scale Modeling - Scholarpedia
    Jan 15, 2010 · Figure 1: Unfiltered and filtered turbulent field showing decomposition used in Large-Eddy Simulation. ... (Leonard 1974). The convolution ...
  64. [64]
    Near-wall treatment for LES models - CFD Online
    Jun 8, 2006 · The most basic form of wall modeling for LES simply imposes some additional constraints upon the eddy viscosity. The standard Smagorinsky ...
  65. [65]
    [PDF] A dynamic subgrid-scale eddy viscosity model
    The model is based on an algebraic identity between the subgrid-scale stresses at two different filtered levels and the resolved turbulent stresses. The subgrid ...
  66. [66]
    [PDF] A proposed modification of the Germano subgrid-scale closure method
    Cabot, and S. Lee, “A dynamic subgrid-scale model for compressible turbulence and scalar transport,” Phys. Fluids. A 3, 2746 (1991). 'P. Moin (private ...
  67. [67]
    (PDF) A scale-dependent dynamic model for large-eddy simulation
    Aug 10, 2025 · A scale-dependent dynamic subgrid-scale model for large-eddy simulation of turbulent flows is proposed. Unlike the traditional dynamic model ...
  68. [68]
    [PDF] Improved subgrid-scale models for large-eddy simulation
    The main objective of this work is to reach a better under- standing of the simplest SGS models in order to derive better models and thereby to improve the.
  69. [69]
    [PDF] An eddy-viscosity subgrid-scale model for turbulent shear flow
    The development of subgrid models for large-eddy simu- lation (LES) is an important area in turbulence research (see the reviews in Refs. 1 and 2). Eddy- ...
  70. [70]
    [PDF] Implicit Large eddy simulation of hypersonic boundary-layer ... - MIT
    In this paper, we perform an implicit large eddy simulation (ILES) of hypersonic boundary-layer transition for a flared cone at Mach number 6.0 and Reynolds ...
  71. [71]
    (PDF) Detached-Eddy Simulation - ResearchGate
    Abstract and Figures. Detached-eddy simulation (DES) was first proposed in 1997 and first used in 1999, so its full history can be surveyed.
  72. [72]
    A New Version of Detached-eddy Simulation, Resistant to ...
    May 30, 2006 · A new version of the technique – referred to as DDES, for Delayed DES – is presented which is based on a simple modification to DES97.
  73. [73]
    (PDF) Improvement of delayed detached-eddy simulation for LES ...
    PDF | Adjustments are proposed of the Delayed Detached Eddy Simulation (DDES) approach to turbulence. They preserve the DDES capabilities particularly.
  74. [74]
    [PDF] Detached Eddy Simulation for Aerospace Applications
    Spalart. “Comments on the Feasibility of LES for Wings, and on Hybrid. RANS/LES Approach”. Proceedings of First AFOSR International Conference on. DNS/LES, 1997 ...
  75. [75]
    [PDF] Application of a Detached Eddy Simulation Approach with Finite ...
    NASA's primary interest in its application to deceleration and maneuvering to enable entry, descent, and landing. ∗. Aerospace Engineer, AIAA Senior Member.
  76. [76]
    [PDF] Evaluation of Detached Eddy Simulation for Turbulent Wake ...
    The focus of this paper is the application of the Detached Eddy Simulation (DES) model to the bluff body wake. DES is perhaps the most popular hybrid RANS/LES ...
  77. [77]
    [PDF] A new version of detached-eddy simulation, resistant to ambiguous ...
    May 30, 2006 · For this purpose, a new version of the technique – referred to as DDES, for Delayed DES – is presented which is based on a sim- ple modification ...
  78. [78]
    (PDF) Recalibrating Delayed Detached-Eddy Simulation to ...
    LES and DES are known to experience the modeled-stress depletion (MSD) phenomenon, resulting in an incorrect application of LES in areas where RANS should be ...
  79. [79]
    Deep structured neural networks for turbulence closure modeling
    Mar 2, 2022 · In the present work, we propose to use a neural network to directly estimate νt, effectively eliminating the need for any additional turbulent ...
  80. [80]
    Artificial neural network approach for turbulence models: A local ...
    The correlation coefficients of the RANS unclosed terms predicted by the LANN model can be made larger than 0.96 in an a priori analysis, and the relative error ...
  81. [81]
    Data-driven discovery of turbulent flow equations using physics ...
    Mar 5, 2024 · In this study, a physics-informed neural network algorithm is utilized to improve the imprecise coefficients of the standard turbulence model ...INTRODUCTION · III. PHYSICS-INFORMED... · Tables related to the relative l2...
  82. [82]
    Simulating Three-dimensional Turbulence with Physics-informed ...
    Jul 11, 2025 · Here we show that appropriately designed PINNs can successfully simulate fully turbulent flows in both two and three dimensions.
  83. [83]
    Using physics-informed enhanced super-resolution generative ...
    This work presents a novel subfilter modeling approach based on a generative adversarial network (GAN), which is trained with unsupervised deep learning (DL)
  84. [84]
    Scientific multi-agent reinforcement learning for wall-models of ...
    Mar 17, 2022 · LES aims to reduce the necessary grid requirements by resolving only the energy-containing eddies and modeling the smaller scale motions.
  85. [85]
    [PDF] AI/ML Applications to Transition and Turbulence Modeling
    Jul 21, 2025 · Duraisamy, K., & Srivastava, V. (2025). Machine learning augmented modeling of turbulence. In Data Driven Analysis and Modeling of Turbulent.
  86. [86]
    [PDF] AN OLD-FASHIONED FRAMEWORK FOR MACHINE LEARNING IN ...
    Aug 1, 2023 · ML is not yet successful in turbulence modeling, with issues like errors in math/physics and overfitting. The field has a mixed reputation and ...
  87. [87]
  88. [88]
    Turbulence closure modeling with machine learning - IOP Science
    This article seeks to review the foundational flow physics to assess the challenges in the context of data-driven approaches.
  89. [89]
    Machine learning in fluid dynamics: A critical assessment - arXiv
    Aug 24, 2025 · The fluid dynamics community has increasingly adopted machine learning to analyze, model, predict, and control a wide range of flows. These ...
  90. [90]
    Generalization Limits of Data-Driven Turbulence Models
    Nov 9, 2024 · In this study, we investigate how data-driven augmentations of popular eddy viscosity models affect their generalization properties.<|control11|><|separator|>