Hostname: page-component-cd9895bd7-mkpzs Total loading time: 0 Render date: 2024-12-26T11:07:01.622Z Has data issue: false hasContentIssue false

Involving non-equilibrium training dataset in data-driven turbulence modeling for turbomachinery

Published online by Cambridge University Press:  28 November 2024

J.L. Du
Affiliation:
Laboratory of Complex Systems, Ecole Centrale de Pékin, Beihang University, Beijing, China
W.S. Liu
Affiliation:
School of Energy and Power Engineering, Beihang University, Beijing, China College of Engineering, Peking University, Beijing, China
L. Fang*
Affiliation:
Laboratory of Complex Systems, Ecole Centrale de Pékin, Beihang University, Beijing, China
T.W. Bao*
Affiliation:
BSS TurboTech, Beijing, China
*
Corresponding authors: L. Fang; Email: [email protected], T.W. Bao; Email: [email protected]
Corresponding authors: L. Fang; Email: [email protected], T.W. Bao; Email: [email protected]
Rights & Permissions [Opens in a new window]

Abstract

In order to improve the performance of $k - \omega $ SST model in turbomachinery, previous studies have used the machine-learning (ML) technique to obtain turbulence models (for example, the ML-RANS EQ model). However, these models do not lead to satisfactory results in complex flows in turbomachinery. In this study, we use non-equilibrium training dataset to obtain a new turbulence model (i.e., the ML-RANS TR-NE-EQ model). Calculations in various cases of turbine cascade flows show that ML-RANS TR-NE-EQ model performs obviously better than ML-RANS EQ model as well as $k - \omega $ SST model.

Type
Research Article
Copyright
© The Author(s), 2024. Published by Cambridge University Press on behalf of Royal Aeronautical Society

Nomenclature

$\boldsymbol{{a}}$

anisotropy Reynolds stress tensor

${C_\epsilon }$

dissipation coefficient

${C_{{p_{{\rm{static}}}}}}$

pressure coefficient

$C_{{p_{{\rm{total}}}}}^{{\rm{loss}}}$

total pressue loss coefficient

$d$

wall distance

${d_b}$

diameter of cylindrical bar in turbine cascades

$h$

channel half width in high-fidelity training data

$i,j$

index

I

unit tensor

$k$

turbulence variable in the $k - \omega $ SST model

$l$

chord in turbine cascades

${l_{ax}}$

axial chord in turbine cascades

${l_b}$

distance of cylindrical bar from the leading edge in turbine cascades

${L_x}$

streamwise location in high-fidelity training data

${\mathcal L}$

integral length scale

$m$

number of layers in the neural network

$n$

number of nodes in the neural network

$N$

number of high-fidelity training data

$p$

pressure

${p_{{\rm{total}}}}$

total pressure

$p_{{\rm{total}}}^{{\rm{inlet}}}$

total pressure at inlet

${q_1}$

first input feature for the practical ML-RANS model

${q_2}$

second input feature for the practical ML-RANS model

${\rm{Re}}$

Reynolds number

${\rm{R}}{{\rm{e}}_m}$

Reynolds number based on bulk mean velocity

${\rm{R}}{{\rm{e}}_\lambda }$

Taylor-scale Reynolds number

${\rm{R}}{{\rm{e}}_\tau }$

Reynolds number based on friction velocity

S

mean strain rate

S h

mean strain rate in high-fidelity training data

${S_k}$

skewness of longitudinal velocity derivative

$t$

pitch of turbine cascades

${u_\tau }$

friction velocity

${U_i}$

component of time-averaged velocity

${U_m}$

bulk mean velocity

${U^ + }$

dimensionless time-averaged velocity

${{\mathcal U}^2}$

turbulent kinetic energy

${w_{i,j}}$

one weight of the neural network

$y$

pitchwise coordinate in turbine cascades

${d^ + }$

dimensionless wall distance

${d^ + }{|_{wall}}$

dimensionless wall distance of the first-level mesh near wall

$z$

Streamwise coordinate in turbine cascades

Greek symbols

${\beta _s},{\beta _1},{\beta _2}$

three kinds of angle in experiment of turbine cascades

${\rm{\Delta }}{x^ + }$

homogeneous streamwise mesh size

${\rm{\Delta }}{y^ + }$

first-level normal mesh size from the wall

${\rm{\Delta }}{z^ + }$

homogeneous spanwise mesh size

$\epsilon $

turbulence dissipation

$\lambda $

L1 regularisation coefficient

$\nu $

kinematic viscosity

${\nu _t}$

eddy viscosity

$\nu _t^m$

optimal eddy viscosity calculated with high-fidelity training data

$\nu _t^{\rm{*}}$

initial learning target of the ML-RANS framework

$\nu _t^0$

target variable for the practical ML-RANS model

$\nu _t^{0,i}$

one target variable for the practical ML-RANS model in high-fidelity training data

$\widehat{{\nu _t^{0,i}}}$

one target variable predicted by the neural network

${\xi _k}$

dimensionless cross diffusion of $k$ and $\omega $

$\boldsymbol{\tau}$

Reynolds stress

$\boldsymbol{\tau}_{h}$

Reynolds stress in high-fidelity training data

$\omega $

turbulence variable in the $k - \omega $ SST model

$\boldsymbol\Omega $

rotation rate

Abbreviations

CFD

computational fluid dynamics

DNS

direct numerical simulation

EQ

equilibrium

FIML

field inversion machine learning

LES

large eddy simulation

M1

method of calculating $\left( {k,\omega } \right)$ with their original physical definition

M2

method of calculating $\left( {k,\omega } \right)$ with $k - \omega $ SST model

ML

machine learning

ML-RANS

iterative machine-learning framework for RANS turbulence modeling

NE

non-equilibrium

PSB

periodic suction and blowing disturbance

RANS

Reynolds-averaged Naviers-Stokes

SST

shear stress transport

TR

transition

1.0 Introduction

In recent decades, numerical simulation methods have been widely applied and have generated a vast amount of flow data. It is still expensive to employ high-fidelity methods such as direct numerical simulation (DNS) and well-resolved large eddy simulation (LES) for complex flows. Thus, the Reynolds-averaged Navier-Stokes (RANS) simulation remains a prevalent method in engineering applications. RANS equations incorporate a Reynolds stress tensor term that necessitates closure through a turbulence model. Towards the end of the 20th century, NASA researchers [Reference Bardina, Huang and Coakley1] conducted a comparative study on popular eddy viscosity models including the $k - \epsilon $ model [Reference Launder and Sharma2], $k - \omega $ model [Reference Wilcox3], $k - \omega $ SST model [Reference Menter4, Reference Menter5], and the Spalart-Allmaras model. They found that the SST model demonstrated the best overall performance because it not only matched the superior performance of other models in simulating simple flows but also excelled in handling complex flows with adverse pressure gradient phenomena. Recently, Chipongo et al. [Reference Chipongo, Khiadani and Sookhak Lari6] compared the simulation outcomes of the realisable $k - \epsilon $ model, $k - \omega $ SST model and the Reynolds stress model (RSM), which directly solves the Reynolds stress transport equations, in spatially varied flows. Taking into account both cost and accuracy, they discovered that these three models were comparably effective. Consequently, this study adopted the SST model as the baseline model.

Despite the prevailing use of the Boussinesq eddy viscosity assumption, there are two issues worthy of attention regarding two-equation models like the SST model. One issue is that deriving turbulence models using theoretical knowledge and observational data derived from equilibrium flows inherently renders these models unsuitable for numerical simulations of non-equilibrium flows at a fundamental level. Fang et al. [Reference Fang, Zhao, Lu, Liu and Yan7] quantitatively described non-equilibrium phenomena in compressor flows using the equation of Lagrangian velocity gradient correlation, revealing that non-equilibrium time-scales are comparable to the main flow time-scales. This underscores the prevalence of non-equilibrium phenomena in turbomachinery. However, as Wilcox [Reference Wilcox8] explains, the scalar attribute and initial definition of turbulence viscosity derive from the rigorous theoretical analysis of the Richardson-Kolmogorov energy cascade. Calibration of turbulence model coefficients heavily relies on decay homogeneous isotropic turbulence and zero-pressure-gradient boundary layers, leading Spalart [Reference Spalart9] to advocate for a reduction in the proportion of cases similar to homogeneous isotropic turbulence, aiming to improve the performance of turbulence models in more practical cases such as jets, wakes, boundary layers, among others.

The other issue is that turbulence variables in eddy viscosity models may not adhere to their initial definitions. As an example, in the $k - \omega $ models, Wilcox [Reference Wilcox8] introduce a stress limiter in the calculation formula for turbulene viscosity, while Menter [Reference Menter4, Reference Menter5] devises a blending function for the SST model. These artificial or empirical enhancements, although improving the predictive accuracy of the $k - \omega $ models, cause the turbulence viscosity and the eddy viscosity model to deviate from its physical underpinnings. Furthermore, the modeled transport equations are not rigorously derived. The closure approximations used in the transport equations for $k$ and $\epsilon $ inevitably reduce the accuracy of the solution, and excessive approximations can even render the transport equations overly dependent on the turbulence data required for modeling. Additionally, the $\omega $ equation is derived via dimensional analysis based on the variation process of specific dissipation rate. Hence, the necessary modeling of the transport equations within eddy viscosity models constitutes a systematic factor contributing to the deviation of turbulence variables from their physical definitions. In Spalart’s perspective [Reference Spalart9], attempting to obtain accurate turbulence variables from turbulence models represents a fallacy in turbulence modeling, given his observation that the best $k - \epsilon $ model arguably possesses an inaccurate $k$ , which significantly complicates the use of DNS data for modeling purposes.

Machine learning (ML) methodologies have been increasingly employed to enhance the predictive capabilities of computational fluid dynamics (CFD), particularly concerning the improvement and modification of established turbulence models [Reference Park and Choi10, Reference Kurz, Offenhäuser and Beck11]. A notable strategy involves leveraging ML techniques to ascertain a generalised formulation of the Reynolds stresses. This stress tensor can be represented as a function of the strain rate ( S) and rotation rate ( $\boldsymbol\Omega $ ), formulated as a linear combination of ten tensorial basis where the coefficients are functions of five scalar invariants [Reference Pope12]. Ling et al. [Reference Ling, Kurzawski and Templeton13] pioneeringly introduced a framework entitled Tensor Basis Neural Networks (TBNN), which maps the local scalar features to the Reynolds stress anisotropy tensor. Following this groundbreaking work, Pope’s seminal analysis [Reference Pope12] has been widely embraced by the majority of research endeavours in the same domain and integrated with diverse methodologies, including the ensemble Kalman method [Reference Zhang, Xiao, Luo and He14], field inversion techniques [Reference Mandler and Weigand15], symbolic regression approaches [Reference Weatheritt and Sandberg16Reference Lav, Banko, Waschkowski, Zhao, Elkins, Eaton and Sandberg18], and random forest algorithms [Reference Wu, Xiao and Paterson19, Reference Brener, Cruz, Macedo and Thompson20], among others.

Another route pursued in the literature involves the application of ML to the transport equations of turbulence models. Notably, Parish and Duraisamy [Reference Parish and Duraisamy21] devised a framework known as field inversion machine learning (FIML). Within this paradigm, the transport equations are adjusted by multiplication with a spatially variant factor inferred through field inversion techniques, while a machine learning model is utilised to deduce regression functions characterising this factor based on selected input features. This methodology has been successfully implemented across various flow regimes, including turbomachinery flows [Reference Ferrero, Iollo and Larocca22], aerofoil aerodynamics [Reference Yan, Li, Zhang and Chen23], boundary layer transition processes [Reference Duraisamy, Zhang and Singh24] and three-dimensional separation flows [Reference Yan, Zhang and Chen25]. Although FIML delivers commendably accurate mean flow fields, it may occasionally produce physically unreasonable Reynolds stresses. Moreover, Zhu et al. [Reference Zhu, Zhang, Kou and Liu26] trained a radial basis function neural network model to replace the conventional Spalart-Allmaras model equations. Concurrently, Zhang et al. [Reference Zhang, Li, You, Kong and Tao27] constructed a novel source term within the $\epsilon $ equation of the $k - \epsilon $ model through deep learning techniques, thereby significantly reducing the overall field error in simulation outputs.

A more cost-effective strategy centers on optimising the turbulence viscosity. Wu et al. [Reference Wu, Xiao and Paterson19, Reference Wu, Xiao, Sun and Wang28] proposed decomposing the anisotropy stress tensor into linear and nonlinear components and computing the optimal eddy viscosity via pointwise least-square approximation. This approach has gained traction among numerous researchers [Reference Liu, Fang, Rolfo, Moulinec and Emerson29Reference Spencer, Przytarski, Adami, Grothe and Wheeler31] and has been empirically shown to enhance the conditioning of RANS equations. More recently, Sun et al. [Reference Sun, Cao, Liu, Zhu and Zhang32] developed an eddy viscosity model, trained on an aerofoil turbulence dataset, which was subsequently tested under high Reynolds number conditions, demonstrating promising results. Liu et al. [Reference Liu, Fang, Rolfo, Moulinec and Emerson29] have introduced an iterative machine learning framework for turbulence modeling, wherein a neural network is trained solely on DNS data of equilibrium channel flow [Reference Abe, Kawamura and Matsuo33, Reference Moser, Kim and Mansour34] to emulate the eddy viscosity computation. This novel approach gives rise to a data-driven turbulence model, denoted as ML-RANS EQ, which demonstrates enhanced performance over the baseline model in scenarios involving separated flow phenomena, such as the two-dimensional periodic hill [Reference Liu, Fang, Rolfo, Moulinec and Emerson29] and the three-dimensional turbine cascade [Reference Fang, Bao, Xu, Zhou, Du and Jin35]. Acknowledging its potential for generalisability, we adopt this framework in the present study, with a introduction provided in sec:theory-1.

Based on Spalart’s statement [Reference Spalart9], turbulence models fail to provide precise estimates of turbulence variables. Our previous work [Reference Fang, Bao, Xu, Zhou, Du and Jin35] illustrates this point, showing that in RANS simulations, the values of $k$ and $\omega $ can significantly differ from reality. In the same work [Reference Fang, Bao, Xu, Zhou, Du and Jin35], we classified the computation methods for turbulence variables in the training phase of machine learning into two categories. One category (M1) involves directly calculating the high-fidelity turbulence variables according to their original physical definitions. The other category (M2) involves solving the transport equations of the turbulence model based on the frozen mean flow field of high-fidelity data. In the prediction phase, all the input features of neural network are constructed by performing the baseline RANS solver. As a result, the M1 method establishes the consistency of the ML model with the DNS field, but does not guarantee consistency with the RANS or LES environment [Reference Duraisamy36], whereas the M2 method ensures the opposite. As pointed out by Wu et al. [Reference Wu, Xiao, Sun and Wang28], even small discrepancies in the Reynolds stress tensor lead to dramatic high error in the mean flow field due to the ill-conditioning nature of RANS equations. Consequnetly, earlier works noted that ML models based on the M1 method often give poor predictions of the mean flow field in a-posteriori tests [Reference Schmelzer, Dwight and Cinnella37Reference Mandler and Weigand39]. Since 1998, Parneix et al. [Reference Parneix, Laurence and Durbin40] introduced the M2 method specifically designed to process and derive turbulence variables from DNS datasets. Amidst the burgeoning field of ML-informed turbulence modeling, several scholars [Reference Liu, Fang, Rolfo, Moulinec and Emerson29, Reference Schmelzer, Dwight and Cinnella37, Reference Weatheritt and Sandberg41, Reference Wu and Zhang42] have advocated for or applied the M2 method to produce turbulence variables critical for machine learning model training. Our preceding research [Reference Fang, Bao, Xu, Zhou, Du and Jin35] compared the turbulence variables based on M1 method and M2 method, and evidenced M2 method’s merits in terms of improved accuracy, leading to its adoption in this investigation.

The aforementioned quantitative investigation by Fang et al. [Reference Fang, Zhao, Lu, Liu and Yan7] has evidenced non-equilibrium phenomena within the regions of corner separation, wake, and boundary layers in compressors. To enhance the precision of simulating such corner separation flows, both velocity helicity reflecting nonlinear turbulence energy backscatter processes [Reference Liu, Lu, Fang and Gao43, Reference Liu, Tang, Scillitoe and Tucker44] and kinematic vorticity accounting for fluid rotation and deformation [Reference Liu, Yan, Fang, Lu, Li and Shao45] have been integrated into the transport equations of conventional turbulence models. Sun [Reference Sun46] showed that employing the SST model augmented with velocity helicity correction, together with a good grid, effectively captures compressor flow characteristics with reasonable fidelity, thereby yielding credible predictions of compressor performance and stage matching. Moreover, Spalart et al. [Reference Spalart9] highlighted that certain canonical flows deviate from the notion of equilibrium conditions. Flows such as normally strained flows, boundary layers experiencing adverse or favourable pressure gradients, steady and unsteady backward-facing step flows, amongst others, have been utilised by Klein et al. [Reference Klein, Craft and Iacovides47] to assess turbulence models’ performance under basic equilibrium regimes. Notably, Liu et al. [Reference Liu, Lu and Fang48] scrutinised and characterised the non-equilibrium turbulence zone within a spatially transitional channel flow. Xiao et al. [Reference Xiao, Wu, Laizet and Duan49] contend that flows over periodically varying hill slopes exemplify strong non-local and non-equilibrium effects, thus conducting direct numerical simulations to establish a database for data-driven turbulence modeling. In complement to embedding non-equilibrium effects within the turbulence modeling framework, as exemplified by Wu and Zhang’s work [Reference Wu and Zhang42], the customisation of training datasets is pivotal to enhancing the efficacy of data-driven turbulence models [Reference Ling, Kurzawski and Templeton13]. For example, incorporating diverse geometries facilitates the generalisability of machine learning (ML)-based turbulence models [Reference Weatheritt and Sandberg16, Reference Wang, Wu and Xiao50]. In this paper, our objective is to engage a training dataset including non-equlibrium flows in data-driven turbulence modeling to refine the accuracy of turbulence models applied in turbomachinery contexts.

The remainder of the paper is organised as follows. In sec:theory, we will recall briefly the theoretical basis of the ML-RANS framework [Reference Liu, Fang, Rolfo, Moulinec and Emerson29, Reference Fang, Bao, Xu, Zhou, Du and Jin35] and non-equilibrium flows. sec:modeling contains details of a new training set and a brief introduction to the training process. The new ML-RANS model has been applied to some in-house cascade test configurations with the previous ML-RANS model and the baseline RANS model. Comparing their simulation results with experimental data, an a-posteriori evaluation of the new ML-RANS model will be presented in sec:application. sec:conclusion is our conclusion and perspectives.

2.0 Theoretical basis

Within this section, we initially present the iterative machine-learning framework for RANS turbulence modeling, herein referred to as ML-RANS, as proposed by Liu et al. previously [Reference Liu, Fang, Rolfo, Moulinec and Emerson29, Reference Liu, Song and Fang51, Reference Liu, Fang, Rolfo, Moulinec and Emerson52]. This framework aims at replacing the conventional calculation formula for turbulence viscosity in the baseline $k - \omega $ SST model using a neural network. Note that all theories and calculations, including a-priori and a-posteriori tests, are 3D. However, statistical average in homogeneous directions can reduce the data dimensions for training and turbulence modeling. Conclusively, we review some research on non-equilibrium turbulence that motivate the incorporation of effects pertaining to three flow regimes within the ML-RANS framework.

2.1 ML-RANS framework

The ML-RANS framework consists of two independent phases: training phase and prediction. Throughout the ML-RANS modeling process, neural networks, input features and target variables play integral roles. Liu et al. [Reference Liu, Fang, Rolfo, Moulinec and Emerson29] employed the TensorFlow platform [Reference Abadi, Barham, Chen, Chen, Davis, Dean, Devin, Ghemawat, Irving, Isard, Kudlur, Levenberg, Monga, Moore, Murray, Steiner, Tucker, Vasudevan, Warden, Wicke, Yu and Zheng53] to construct an artificial neural network. In the training phase, M2 method described in 1 is first adopted for the construction of machine learning dataset. The machine learning regression algorithms, along with prescribed input features and target variables, are used to determine the weights and biases of the neural network, resulting in a machine learning model. During the prediction phase, this machine learning model is coupled with a baseline turbulence model. It first computes the input features using numerical solutions of both transport equations in the baseline turbulence model and RANS equations. The comparison of turbulent quantities ( $k$ , $\omega $ , …) obtained with DNS/LES with those of RANS formulation can refer to previous studies, for example Figs. 1 and 2 in Ref. [Reference Fang, Bao, Xu, Zhou, Du and Jin35]. Subsequently, these input features are mapped onto the target variable to generate turbulence viscosity. This framework exhibits two notable attributes: consistency in the computation of input features across both phases and the evolution of the turbulence viscosity within the prediction phase.

Figure 1. Comparisons of $\left( {{q_1},{q_2}} \right)$ in the transitional channel flow in transition region (TR), non-equilibrium region (NE) and equilibrium region (EQ) respectively [Reference Fang, Bao, Xu, Zhou, Du and Jin35].

In the ML-RANS framework, the target variable refers to the object of turbulence modeling, which involves leveraging machine learning techniques to repesent Reynolds stresses, either as mathematical expressions or through artificial neural networks, derived from high-fidelity flow data. However, certain researchers [Reference Thompson, Sampaio, de Bragança Alves, Thais and Mompean54, Reference Poroseva, Colmenares and Murman55] have observed discrepancies between the resolved mean velocity when employing the Reynolds stresses from high-fidelity data to close RANS equations and the mean velocity in the original high-fidelity data. Wu et al. [Reference Wu, Xiao, Sun and Wang28] highlighted that the RANS equations coupled with Reynolds stress closure models could be ill-conditioned, where minor a-priori errors in modeled terms, such as the Reynolds stresses, can escalate into significant a-posteriori errors during the solution of the RANS equations. To tackle this issue, Wu et al. [Reference Wu, Xiao, Sun and Wang28] partitioned the Reynolds stresses into an implicitly treatment (or linear) component involving ${\nu _t}$ and a nonlinear part, i.e.,

(1) \begin{align}\boldsymbol{{a}} = \boldsymbol\tau - \frac{1}{3}tr( \boldsymbol{\tau})\boldsymbol{{I}} = 2{\nu _t}\boldsymbol{{S}} + {\boldsymbol{{a}}_ \bot }.\end{align}

The linear portion corresponds to one aspect of Pope’s proposed mathematical representation of the Reynolds stress [Reference Pope12], which adheres to the Boussinesq eddy viscosity hypothesis. From the high-fidelity dataset, the Reynolds stress tensor τ h and the mean shear stress tensor S h are extracted to calculate ${\nu _t}$ by minimising the nonlinear part, resulting in

(2) \begin{align}{\nu _t} = \left| {\frac{{{\tau _{{h_{ij}}}}{S_{{h_{ij}}}}}}{{2{S_{{h_{ij}}}}{S_{{h_{ij}}}}}}} \right|.\end{align}

Their analysis showed that enhancing the implicit treatment of the Reynolds stress, for instance, by adding an increment to the optimal ${\nu _t}$ while correspondingly subtracting from the nonlinear part, can further improve the conditioning of the model, thereby reducing the sensitivity of the RANS equation solution to the unclosed terms [Reference Wu, Xiao, Sun and Wang28]. In earlier work, Wu et al. [Reference Wu, Xiao and Paterson19] had independently trained machine learning models for the linear and nonlinear components of the Reynolds stress, culminating in a well-conditioned and accurate Reynolds stress model. Liu et al. [Reference Liu, Fang, Rolfo, Moulinec and Emerson29, Reference Liu, Song and Fang51, Reference Liu, Fang, Rolfo, Moulinec and Emerson52], building upon this, took a step further by focusing solely on the turbulence viscosity as the target variable, bypassing the nonlinear part of the Reynolds stress altogether. To enhance the effectiveness of the machine learning regression, they first nondimensionalised the turbulence viscosity using the traditional definition $k/\omega $ from the $k - \omega $ model [Reference Wilcox3, Reference Wilcox8], followed by normalisation to arrive at the actual form of the target variable presented in Table 1.

Table 1. The input features and the target variable for the practical ML-RANS model [Reference Liu, Song and Fang51, Reference Liu, Fang, Rolfo, Moulinec and Emerson52]

Figure 2. Mean velocity profiles by use of different high-fidelity methods.

As for the input features, Liu et al. [Reference Liu, Fang, Rolfo, Moulinec and Emerson29] initially derived from Pope’s analysis [Reference Pope12] an essential input feature (i.e., $k/\nu \omega $ ) related to the target variable. Based on empirical evidence, Liu et al. bypassed tensor invariances in the input features to stabilise simulations and introduced five supplementary input features that included turbulence intensity, wall distance, cross-diffusion of $k$ and $\omega $ , along with two SST model variables specifically capturing properties of the viscous sublayer and the turbulent region [Reference Liu, Fang, Rolfo, Moulinec and Emerson29]. In subsequent research [Reference Liu, Song and Fang51, Reference Liu, Fang, Rolfo, Moulinec and Emerson52], they streamlined their approach by retaining only the first two input features and subjected them to normalisation based on their initial design, as detailed in Table 1. Notably, our previous study [Reference Fang, Bao, Xu, Zhou, Du and Jin35] has demonstrated the ability to identify distinct regimes of turbulence development exclusively using these two dimensionless and normalised input features $\left( {{q_1},{q_2}} \right)$ , as illustrated in Fig. 1. In previous practices, DNS datasets [Reference Abe, Kawamura and Matsuo33, Reference Moser, Kim and Mansour34] and RANS calculations [Reference Liu, Fang, Rolfo, Moulinec and Emerson29, Reference Liu, Song and Fang51, Reference Liu, Fang, Rolfo, Moulinec and Emerson52] were used, respectively.

2.2 Non-equilibrium flows

The property of multi-scale energy transfer is a good starting point for understanding turbulence. The well-known Richardson-Kolmogorov energy cascade indicates an equilibrium dissipation law, that can be depicted as a balance of high Reynolds turbulence between the forward energy transfer and the dissipation.

(3) \begin{align}\epsilon = \frac{{{C_\epsilon }{{\mathcal U}^3}}}{{\mathcal L}}.\end{align}

where ${{\mathcal U}^2}$ is turbulent kinetic energy, ${\mathcal L}$ is the integral length scale and the dissipation coefficient ${C_\epsilon }$ is constant. In fact, in real flows there exist various dissipation laws, including the equilibrium or quasi-equilibrium laws and the non-equilibrium laws, as Vassilicos [Reference Vassilicos56] introduced. It was discovered that there exists a non-equilibrium dissipation law between ${C_\epsilon }$ and Taylor-scale Reynolds number ${\rm{Re}}_{\lambda}$ , which is approximately ${C_\epsilon }\sim {\rm{Re}}_\lambda ^{ - 1}$ , or rigorously by theoretical analysis [Reference Bos and Rubinstein57],

(4) \begin{align}{C_\epsilon }\sim {\rm{Re}}_\lambda ^{ - \frac{{15}}{{14}}}.\end{align}

It denotes the imbalance between energy production and dissipation, which mainly exists with a spatially inhomogeneous geometry, even when the flow is a fully developed turbulence. The non-equilibrium canonical flows discussed in sec:intro largely conform to this dissipation rate. Researchers have actively explored the integration of non-equilibrium phenomena within turbulence models by leveraging the ratio between the production and dissipation of turbulence kinetic energy [Reference Wu and Zhang42, Reference Li, Zhang and Chen58]. Recently, it was found that for flows whose transfer spectrum is out-of-equilibrium, there also exists another non-equilibrium dissipation law

(5) \begin{align}{C_\epsilon }\sim {\rm{Re}}_\lambda ^{ - 2},\end{align}

which is usually rather evident in short-time evolution by comparing to the $ - 15/14$ scaling [Reference Liu, Lu, Bos and Fang59Reference Araki and Bos62] and is closely related to the self-organisation procedure in a flow that is (temporally or spacially) not fully developed. During the evolution of turbulent flows, this novel scaling law typically implies the presence of turbulence energy backscatter (i.e., the transfer of energy from smaller to larger scales). Despite their appearance prior to the establishment of such a scaling law, the conventional turbulence modeling efforts [Reference Liu, Lu, Fang and Gao43, Reference Liu, Tang, Scillitoe and Tucker44] described in sec:intro, aiming at enhancing the Spalart-Allmaras and SST models by incorporating helicity, can be regarded as early attempts to account for this out-of-equilibrium in third-order statistical quantities. However, until now, there has not yet been a data-driven turbulence modeling that specifically addresses the $ - 2$ scaling behaviour associated with energy backscatter phenomena.

In previous studies we showed that a spatially transitional channel flow can perfectly involve all the known states of energy transfer of turbulence evolution, including both the $ - 15/14$ and $ - 2$ scalings of non-equilibrium flows [Reference Liu, Lu, Bos and Fang59]. On another hand, we also illustrated that the flows in turbomachinery are almost non-equilibrium everywhere [Reference Fang, Zhao, Lu, Liu and Yan7]. In the present paper we then use this dataset, aiming at involving various flows in the training procedure, and hoping to improve the performance of turbulence models in turbomachinery.

3.0 Turbulence modeling

According to earlier research [Reference Weatheritt and Sandberg16, Reference Wang, Wu and Xiao50], the test flow’s shape, pressure gradient, and other components should be present in the training flow as well. Although its training set was created using the canonical channel flow, the ML-RANS EQ model performed slightly better than its baseline model in the complex geometry (see more details in the papers of Liu et al. [Reference Liu, Fang, Rolfo, Moulinec and Emerson29] and Fang et al. [Reference Fang, Bao, Xu, Zhou, Du and Jin35]), indicating good generalisation of the present ML-RANS framework. Using a spatially transitional channel flow database [Reference Liu, Lu and Fang48] with transition region (TR), non-equilibrium region (NE), and equilibrium region (EQ), we discovered three essential constraints of this training set (i.e., sampling in one dimension, geometry of the straight wall, and turbulence in equilibrium) in the previous study [Reference Fang, Bao, Xu, Zhou, Du and Jin35]. Moreover, we specified the issue of expansibility from EQ to TR/NE in both a-priori and a-posteriori analysis. In the present study, we create the ML-RANS TR-NE-EQ model by fusing the aforementioned transitional flow database [Reference Liu, Lu and Fang48] with the present ML-RANS framework [Reference Liu, Song and Fang51, Reference Liu, Fang, Rolfo, Moulinec and Emerson52].

3.1 Training set

A training set based on the spatially transitional channel flow [Reference Liu, Lu and Fang48] will be introduced and evaluated in this subsection. The channel has a $64{\pi}h \times 2h \times 2{\pi}h$ domain with $h$ channel half-width with a spatial resolution $1,024 \times 96 \times 64$ in $x$ , $y$ , $z$ directions, including a PSB (periodic suction and blowing disturbance) region in the streamwise interval ${L_x} \in \left[ {{\pi}h,2{\pi}h} \right]$ and a fringe region [Reference Ma, Van Doorne, Zhang and Nieuwstadt63] in the streamwise interval ${L_x} \in \left[ {56{\pi}h,64{\pi}h} \right]$ . The existence of two regions contributes to the generation of the spatial transition. The commonly used dynamic Smagorinsky model was selected as the subgrid-scale model for LES of the channel flow. No-slip boundary conditions are imposed at wall (excluding the PSB region), whereas periodic boundary conditions are utilised along the spanwise direction with the LES. Laminar flow distribution is set at inlet and zero-gradient is set at outlet. Furthermore, the grid spacings and Reynolds numbers are documented in Table 2 for more details. More details refer to Refs [Reference Liu, Lu and Fang48, Reference Xu, Zhang, den Toonder and Nieuwstadt64, Reference Liu65].

Table 2. Information of LES on the spatially transitional channel flow [Reference Liu, Lu and Fang48]

The mean flow profile deviates from its parabolic shape within the range ${L_x} \gt 12\pi $ with the maximum value of the mean velocity reaching approximately 1.14 at ${L_x} \approx 23.4\pi $ . This confirms the typical morphology of a turbulent velocity profile beyond ${L_x} \gt 23.4\pi $ , which is in good agreement with DNS results from Moser et al. [Reference Moser, Kim and Mansour34] and those from Abe et al. [Reference Abe, Kawamura and Matsuo33] at the same friction Reynolds number, ${\rm{R}}{{\rm{e}}_\tau }$ . However, the flow had not yet been fully developed until nearly ${L_x} = 31\pi $ , at which the profile of the total shear stress was a horizontal straight line. In the interval, ${L_x} \in \left[ {23.4{\pi}h,31{\pi}h} \right]$ , the dissipation coefficient ${C_\epsilon }$ increased weakly to constant, and the skewness of longitudinal velocity derivative ${S_k}$ decreased slightly to constant. In addition, ${C_\epsilon }$ was inversely proportional to the integral scale Reynolds number at a low Reynolds number. The phenomena were also observed in the non-equilibrium study of grid turbulence [Reference Hearst and Lavoie66]. Consequently, the segment $\left[ {23.4{\pi}h,31{\pi}h} \right]$ along the streamwise direction was classified as NE, while ${L_x} \lt 23.4\pi $ was characterised as TR and ${L_x} \gt 31\pi $ was considered as EQ. The near-wall mean velocity profile at EQ phase is compared with DNS results of Abe et al. [Reference Abe, Kawamura and Matsuo33], as well as the experiment of Hussain and Reynolds [Reference Hussain and Reynolds67]. Pope reviewed the classical wall law, including the linear relation ${U^ + } = {d^ + }$ in the viscous sublayer ( ${d^ + } \lt 5$ ) and the logarithmic law

(6) \begin{align}{U^ + } = \frac{1}{\kappa }{\rm{ln}}({d^+}) + C,\end{align}

due to von Karman for ${d^ + } \gt 30$ , where $\kappa \approx 0.41$ represents the Karman constant and $C$ was taken as 5.2 despite dependance on Reynolds number [Reference Abe, Kawamura and Matsuo33, Reference Pope68]. As shown in Fig. 2, all three simulations and the experiment agree with the linear relation in the viscous sublayer. In the logarithmic region, the DNS result at ${\rm{R}}{{\rm{e}}_\tau } = 640$ closely matches both the theoretical formula and corresponding experimental findings at the same Reynolds number, thereby substantiating the accuracy of the DNS programme employed by Abe et al. At ${\rm{R}}{{\rm{e}}_\tau } = 395$ , the LES result aligns closely with the DNS result, which indicates that the LES result remains highly credible. For in-depth information, see the paper published by Liu et al. [Reference Liu, Lu and Fang48].

In TR, positive ${S_k}$ denotes the energy backscatter signal [Reference Zhao, Liu, Shao, Fang and Dong69], which also occurs after altering the sign of velocities in time-reversed turbulence [Reference Liu, Lu, Bos and Fang59] where the large-scale dynamics cannot feel the effect of energy backscatter as concurrently as small-scale dynamics does. This non-equilibrium behaviour, characterised by a $ - 2$ scaling law, continues to impact the channel flow until the large-scale structures within NE are notably affected. Subsequently, the flows begins to conform to $ - 15/14$ scaling law. The flow eventually transforms into equilibrium turbulence in EQ after the rapid and slow non-equilibrium procedures. The physical information of transition, non-equilibrium and equilibrium turbulence is retrieved from the flow data in the two-dimensional mid-span section of the spatially transitional channel flow in the current study.

3.2 Modeling process

During the training phase, Google’s TensorFlow [Reference Abadi, Barham, Chen, Chen, Davis, Dean, Devin, Ghemawat, Irving, Isard, Kudlur, Levenberg, Monga, Moore, Murray, Steiner, Tucker, Vasudevan, Warden, Wicke, Yu and Zheng53] was adopted as the platform for machine learning. The training dataset was separated randomly into the training part and the validation one in a ratio of four to one. According to training effect, the hyperparameters in Table 3 were adopted to construct a feedforward artificial neural network (i.e., multi-layer perception). For sake of prevention from overfitting [Reference Tibshirani70], L1 regularisation was adopted to minimise the weight of each layer in the neural network. The total loss was thus defined as

(7) \begin{align}{\rm{Loss}} = \frac{1}{N}\mathop \sum \limits_{i = 1}^N {\left(\widehat{\nu _t^{0,i}} - \nu _t^{0,i}\right)^2} + \lambda \mathop \sum \limits_{i,j}^{m,n} | {{w_{i,j}}} |.\end{align}

where $\nu _t^{0,i}$ is the target variable in the training data, $\widehat{\nu _t^{0,i}}$ is the value predicted by the neural network, and ${w_{i,j}}$ is the weight in the neural network. $\lambda $ represents the L1 regularisation coefficient. $N$ is the number of training data, $m$ and $n$ are respectively the numbers of layers and nodes in the neural network. The first part of the loss is the mean square error (MSE) from the training data, and the second part is the L1 regularisation penalty. As seen in Fig. 3, the evolution of errors indicates that our model is rapidly established once the training process starts. To estimate visually the model result, we plot the profiles of the target variable $\nu _t^0$ at different streamwise locations in Fig. 4, and the relation between the target variable and both input features is shown in Fig. 5. It can be observed that the training data and ML predictions are all in good agreement.

Table 3. Hyperparameters of the neural network for the ML-RANS TR-NE-EQ model

Figure 3. Evolution of loss in the training process. The x-axis represents the ratio of current training step and total steps.

Figure 4. Profiles of the target variable $\nu _t^0$ at $x/( {{\pi}h} ) = 20,30,40,50$ in TR (blue), NE (red) and EQ (green) regions: comparison of ML predictions (lines) and a snapshot of training dataset (points).

Figure 5. Relation between the target variable $\nu _t^0$ and two input features $\left( {{q_1},{q_2}} \right)$ at four different streamwise locations: comparison of ML predictions (line) and a snapshot of training dataset (point).

Following training, a comprehensive data-driven turbulence model is created by combining the machine learning model with the predictor library [Reference Liu, Song and Fang51]. Figure 6 illustrates that the target variable $\nu _t^0$ is pretty independent of the input features where ${q_1}$ is large for the ML-RANS EQ model, while it decreases obviously as ${q_2}$ increases in the same region for the ML-RANS TR-NE-EQ model.

Figure 6. Mapping from $\left( {{q_1},{q_2}} \right)$ to $\nu _t^0$ in two ML-RANS models.

3.3 A-posteriori test

Having implemented the novel ML-RANS TR-NE-EQ model in OpenFOAM, we applied it into the spatially transitional channel flow used during training, comparing its performance against the earlier ML-RANS EQ model and the baseline model. Inspired by Zhao’s doctoral dissertation [Reference Zhao71] where several turbulence models were utilised for RANS simulations of this transitional channel flow, this study followed suit by removing the fringe area close to the outlet. A computationally economical mesh configuration of $440 \times 50 \times 32$ was established across the computational domain of $55{\pi}h \times 2h \times 2{\pi}h$ , ensuring effective refinement near the top and bottom walls, particularly with the first cell layer meeting the requirement of ${d^ + } \lt 1$ .

Consistent with LES conducted by Liu et al. [Reference Liu, Lu and Fang48], the transition is facilitated by the PSB region near the inlet; thus, for the wall segment where the streamwise coordinates ${L_x}$ lie within $\left[ {{\pi}h,2{\pi}h} \right]$ , we also designed the PSB boundary condition for the velocity field while imposing a zero-gradient condition for other physical quantities. For the remaining walls, we set the no-slip condition with low-Reynolds number corrections. In the meantime, laminar flow inlet and zero-gradient outlet were configured accordingly. Considering that the velocity boundary values vary over time in the PSB region, we conducted unsteady RANS simulations using the baseline turbulence model, ML-RANS EQ model, and the ML-RANS TR-NE-EQ model, each assisted by the PISO algorithm. To maintain numerical stability, the time step was constrained to ensure a Courant number below 1.

For incompressible flows, the bulk mean velocity remains constant, and the transition from laminar to turbulence can be reflected through variations in the maximum value of the mean velocity. Upon comparing the simulation outputs generated by the three turbulence models against the maximum value of the mean velocity from the LES data, it is observed that the ML-RANS TR-NE-EQ model delivers superior performance among three RANS turbulence models, as shown in Fig. 7. Firstly, both ML-RANS models effectively avoid the occurrence of the local velocity peak within the PSB region simulated by the baseline model. Subsequently, following the PSB region, the maximum value of the mean velocity computed by all three models begin to decline, which indicates that the flow starts to deviate from the laminar. However, the ML-RANS TR-NE-EQ model exhibits a notably delayed initiation of flow transition, as its calculated mean velocity maximum doesn’t experience a sharp drop until reaching ${L_x} = 10{\pi}h$ , aligning closely with the LES-derived transitional point ${L_x} = 12{\pi}h$ . The RANS simulation adopting the ML-RANS TR-NE-EQ model presents remarkable agreement with the LES results in the NE region. Within the EQ region where the large-scale non-equilibrium is virtually absent, the maximum value of the mean velocity calculated by all four numerical simulations converge. Among three RANS models, the ML-RANS TR-NE-EQ model’s prediction shows a closer resemblance to the LES result than that of the baseline model. In summary, the ML-RANS TR-NE-EQ model provides the best simulation of the transitional channel flow among the three RANS turbulence models.

Figure 7. Evolution of maximum value of the mean velocity along streamwise direction.

4.0 Application in the turbine cascades

We apply two practical ML-RANS models and their baseline turbulence model (i.e., $k - \omega $ SST model) to some turbine cascade flow cases which had been the subject of in-house experimental measurements, respectively. Every cascade in the experiment consists of the same eight blades, which stand in for the mid-span region of a low-pressure turbine rotor blade. Several cylindrical bars are fitted as wake generators in front of these blades. The total pressure, total temperature and wall static pressure in the cascade inlet are all monitored for the flow conditions. Certain aerodynamic parameters are monitored using an L-shaped five-hole probe and wall static pressure holes at a distance of $0.4$ axial chord ( ${l_{ax}}$ ) from the blade’s trailing edge (e.g., total pressure loss coefficient and Mach number). In the present study, we analyse 15 cases with five various blade profiles (denoted as Blade 1, Blade 2, …, respectively) and three different angles of attack ( ${0^ \circ }, - {15^ \circ }$ , and ${15^ \circ }$ , respectively), under similar outlet Reynolds number (i.e., 770,000) and outlet Mach number (i.e., 0.75). Figure 8 illustrates a few cascade case configurations by using randomly Blade 2 and zero angle-of-attack as an example. In the figure the star symbols located $0.4{l_{ax}}$ before the leading edge are marked, corresponding to the observation positions of spanwise near-wall behaviour in simulation, which will be represented later. More experimental details can refer to our previous study [Reference Fang, Bao, Xu, Zhou, Du and Jin35].

Figure 8. Sketch of cascade design of LPT: near-wall behaviours in simulations are observed both at the blue star (endwall, measurement along the $x$ axis) and at red arrows (blade).

All the computational domains containing five different blades are divided by an H–O–H structured mesh of 193 points (for Blade 3) or 185 points (for other four blades) in the streamwise direction and 84 points in the orthogonal direction, with the first node near the blade satisfying ${d^ + }{|_{wall}} \lt 1$ . The length of the inlet region is half of the axial chord length, while the outlet region has one axial chord length. Figure 9 shows the mesh about Blade 2. Mesh independency has been accordingly verified to guaranty the rationality of simulation. The $k - \omega $ SST model and two ML-RANS models are applied in the open-source code OpenFOAM. In practice, we set uniform values of total pressure, total temperature and flow angle at the inlet, a static pressure at the outlet, and no-slip condition at the wall with low Reynolds number corrections.

Figure 9. Mesh of the channel with Blade 2 used in numerical simulations.

The comparison on near-wall effects between ML-RANS models and the classic $k - \omega $ SST model is shown in Figs. 10 and 11. In front of bars and blades, the endwall of the turbine cascade resembles a channel. The near-wall results observed here are not influenced by the wake. The mean velocity profile in Fig. 10(a) by using ML-RANS TR-NE-EQ model (i.e., ${U^ + } = 2.46{\rm{ln}}\!({d^+}) + 5.92$ ) is the closest to the logarithmic law of the wall (see Equation 6), by comparing to the other two curves. In addition, the eddy viscosity shown in Fig. 10(b) also differs. At locations on the blade surface away from the endwalls, two measurement points were selected near the leading edge and trailing edge, respectively, to compare the mean velocity profiles along the blade surface normal direction from different models. Given that molecular viscosity dominates the flow within the viscous sublayer, the RANS equations necessarily simplify to the linear relation, as calculated by all three models. In the logarithmic region, however, there were slight differences in performance among the three models near both the leading and trailing edges of the blade. Similar to Fig. 10, the $k - \omega $ SST model and the ML-RANS EQ model provided distinct logarithmic laws that deviated from the von Kármán’s formula. In contrast, the ML-RANS TR-NE-EQ model approximated the standard logarithmic law closely near the leading edge but showed a obvious difference only in the constant term $C$ of the Equation (6) near the trailing edge.

Figure 10. Comparison on results near endwall among the $k - \omega $ SST model and the two ML-RANS approaches, at the location marked as the blue star in Fig. 8: (a) Mean velocity, (b) Eddy viscosity.

Figure 11. Comparison on mean velocities near blade among the $k - \omega $ SST model and the two ML-RANS approaches, at the locations marked as the arrows with different distances from the leading edge in Fig. 8: (a) $0.2{l_{ax}}$ , (b) $0.7{l_{ax}}$ .

A-posteriori results in Fig. 12 show that both the two ML-RANS approaches show good predictions for the static pressure coefficient (Equation 8) around the blade. In the previous study [Reference Fang, Bao, Xu, Zhou, Du and Jin35], we pointed out that the blade downstream without separation is almost EQ according to Fang et al. [Reference Fang, Zhao, Lu, Liu and Yan7] and that the bar wake with obvious separation is mostly NE/TR according to Tao and Zhou [Reference Tao and Zhou72]. We take the case of Blade 2 and zero angle-of-attack as an example. We comment that this choice does not mean that Blade 2 is the best case. The global performance of all cases will be shown later. The Mach number calculated by the ML-RANS TR-NE-EQ model in Fig. 13 is the smallest among three simulations both in the bar wake and the blade downstream, indicating that its bar wake is the longest and that its blade downstream is the most obvious, while the ML-RANS EQ model yields the similar effect to those of the $k - \omega $ SST model. Eddy viscosity is crucial for solving the Reynolds equation in the eddy viscosity model. When comparing the scalar in the bar wake, it can be observed that the ML-RANS TR-NE-EQ model exports far less eddy viscosity than the ML-RANS EQ model and their baseline model (see Fig. 14).

Figure 12. Comparison on $\mathrm{a}$ - $\mathrm{posteriori}$ results on static pressure coefficient around the blade.

Figure 13. Mach number (Ma) calculated by (a) ML-RANS TR-NE-EQ model, (b) $k - \omega $ SST model and (c) ML-RANS EQ model at mid-span section: 0.3 at cyan contour and 0.7 at yellow contour.

Figure 14. Eddy viscosity ( ${\nu _t}$ ) calculated by (a) ML-RANS TR-NE-EQ model, (b) $k - \omega $ SST model and (c) ML-RANS EQ model at at mid-span section.

Furthermore, the ML-RANS EQ and the baseline model both depend on equilibrium turbulence and two ML-RANS models were born in the same data-driven turbulence modeling framework [Reference Liu, Fang, Rolfo, Moulinec and Emerson29], where the machine learning regression mapping partially replaces the physics-informed equation of ${\nu _t}$ . Thus, analysing discrepancies in turbulence viscosity outputs from two ML-RANS models can help us understand influence of non-equilibrium flows on turbulence modeling. One practical strategy involves indicating the TR, NE and EQ regimes, especially the TR/NE zones. Nevertheless, based on prior research [Reference Fang, Zhao, Lu, Liu and Yan7, Reference Fang, Bao, Xu, Zhou, Du and Jin35, Reference Liu, Lu and Fang48, Reference Fang and Bos60], the quantitative criterion needs further determination to accurately identify non-equilibrium flow state Inspired by the successful partition of three regions, the $\left( {{q_1},{q_2}} \right)$ diagram is utilised to identify the discrepancy of two ML-RANS models in the bar wake satisfying ${\nu _t} \gt 0.0015{{\rm{m}}^2}/{\rm{s}}$ . Two ML-RANS models in Fig. 15 predicate that the closer the position is to the bar, the greater the couple $\left( {{q_1},{q_2}} \right)$ , which denotes that the bar wake is located in TR/NE regimes according to Fig. 1. Reviewing Fig. 6, due to diversity of training data, the ML-RANS TR-NE-EQ model predicts more changeful target variable than ML-RANS EQ model in TR/NE regions, which indicates that the new model can deal with different flow phases without requiring manual adjustments. As a result, the ML-RANS TR-NE-EQ model offers lower turbulence viscosity than two other models in the bar wake. Note that these results are predicted by extrapolation instead of interpolation. Usually the extrapolation of ML is poor, but it is unavoidable when we employ ML-RANS models to real engineering applications, in particular for those research and development projects considering off-design conditions, which has attracted gradually more attentions in previous years. Thus we believe that the present attempt by using extrapolation is important and necessary for real engineering applications.

(8) \begin{align}{C_{{p_{{\rm{static}}}}}} = \frac{p}{{p_{{\rm{total}}}^{{\rm{inlet}}}}},\end{align}

Figure 15. Mapping from $\left( {{q_1},{q_2}} \right)$ to $\nu _t^0$ of two ML-RANS models in the bar wake ( ${\nu _t} \gt 0.0015{{\rm{m}}^2}/{\rm{s}}$ ).

Figure 16. Total pressure loss coefficient at measurement plane on the blade mid-span section.

Figure 17. Mach number at measurement plane on the blade mid-span section.

Figure 18. Deviation of the second peak about the total pressure loss coefficient and the Mach number. The solid and larger symbols represent the result in the selected case (Blade 2 with zero angle-of-attack).

In order to evaluate the ML-RANS TR-NE-EQ model, we compare the simulation results of various turbulence models with the corresponding experimental results in the blade downstream. Figure 16 illustrates the difference in the total pressure loss coefficients (Equation 9). In the five simulation results, the first peak (i.e., $y \approx 0.2t$ ) in the y axis direction in the figure corresponds to the wake of the bar, whereas the second peak (i.e., $y \approx 0.5t$ ) corresponds to the wake of the blade. Because of the substantial interaction between the bar wake and the blade pressure side boundary layer, the first peak (i.e., $y \approx 0.35t$ ) in the experiment is closer to the second peak and is less noticeable. Such a phenomenon is captured by neither the conventional nor the ML-RANS model. However, the ML-RANS TR-NE-EQ model predictes rather accurately the total pressure loss in the blade wake, while the ML-RANS EQ model offers a modest improvement. The aforementioned phenomena are also observed in the prediction for the Mach number (see Fig. 17) as long as we pay attention to the valley instead of the peak. For each of the 15 cases, we continue to observe the discrepancy between the experimental result and the simulation results from various turbulence models on the measure plane behind the cascade. Equation (10) is utilised to calculate the deviation of the second peak (i.e., the blade wake) about the total pressure loss coefficient and the second valley (i.e., the blade wake) about the Mach number. The ML-RANS TR-NE-EQ model performs better than others, in particular captures well the valley of Mach number in the blade wake (see Fig. 18). The solid and larger symbols represent the result in the selected case, i.e., Blade 2 with zero angle-of-attack, but for other cases the trend is exactly similar: the ML-RANS TR-NE-EQ model performs better than other models, in the sense of a global consideration on both Mach number and pressure loss coefficient.

(9) \begin{align}C_{{p_{{\rm{total}}}}}^{{\rm{loss}}} = \frac{{p_{{\rm{total}}}^{{\rm{inlet}}} - {p_{{\rm{total}}}}}}{{{p_{{\rm{total}}}} - p}},\end{align}
(10) \begin{align}{\rm{Deviation}} = \frac{{\left| {{\rm{Simulation\;\;result}} - {\rm{Experimental\;\;result}}} \right|}}{{\left| {{\rm{Simulation\;\;result}}} \right| + \left| {{\rm{Experimental\;\;result}}} \right|}}.\end{align}

In consequence, the ML-RANS TR-NE-EQ model cannot only identify the TR/NE in the bar wake with $\left( {{q_1},{q_2}} \right)$ like the ML-RANS EQ model but also transform them to the significantly different eddy viscosity from one calculated by the equilibrium model. The discrepancy is then transferred to the blade wake on account of the interaction between the bar wake and the boundary layer in the blade pressure side. The model improvement is in particular obvious in the region of blade wake by comparing to experiment.

5.0 Conclusions

In previous studies, the ML-RANS framework has presented better prediction for flow statistics such as the total pressure loss coefficient downstream of the blade [Reference Liu, Fang, Rolfo, Moulinec and Emerson29, Reference Fang, Bao, Xu, Zhou, Du and Jin35]. However, results were still quite different comparing to experiments, which might stem from the fact that these ML-RANS models were trained by using equilibrium turbulent flow dataset (i.e., ML-RANS EQ model). Considering the non-negligible non-equilibriurm effect of energy transfer in real turbomachinery [Reference Fang, Zhao, Lu, Liu and Yan7], in the present paper we then aim at improving the ML-RANS models by involving non-equilibrium training dataset (i.e., ML-RANS TR-NE-EQ model). Comparing with ML-RANS EQ model and the baseline RANS model (i.e., $k - \omega $ SST model), we show that ML-RANS TR-NE-EQ model with non-equilibrium features predicts better statistics, and shows good agreement with experiments.

Acknowledgements

We would like to express our gratitude to O. Jung and J. Pavillet for their supportive guide to the project. This work is supported by National Natural Science Foundation of China (Project Approval No. 12372214, U2341231), BSS Project (PO_20201112_BUAA PhD Project) and the Science Center for Gas Turbine Project (Project No. P2022-C-III-001-001).

References

Bardina, J., Huang, P. and Coakley, T. Turbulence modeling validation, 28th Fluid Dynamics Conference, American Institute of Aeronautics and Astronautics, June 1997.CrossRefGoogle Scholar
Launder, B.E. and Sharma, B.I. Application of the energy-dissipation model of turbulence to the calculation of flow near a spinning disc, Lett. Heat Mass Transfer, 1974, 1, (2), pp 131137.CrossRefGoogle Scholar
Wilcox, D.C. Reassessment of the scale-determining equation for advanced turbulence models, AIAA J., 1988, 26, (11), pp 12991310.CrossRefGoogle Scholar
Menter, F.R. Zonal two equation k-w turbulence models for aerodynamic flows, 23rd Fluid Dynamics, Plasmadynamics and Lasers Conference, American Institute of Aeronautics and Astronautics, 1993.CrossRefGoogle Scholar
Menter, F.R. Two-equation eddy-viscosity turbulence models for engineering applications, AIAA J., 1994, 32, (8), pp 15981605.CrossRefGoogle Scholar
Chipongo, K., Khiadani, M. and Sookhak Lari, K. Comparison and verification of turbulence Reynolds-averaged Navier–Stokes closures to model spatially varied flows, Sci. Rep., 2020, 10, (1), p 19059.CrossRefGoogle ScholarPubMed
Fang, L., Zhao, H.K., Lu, L.P., Liu, Y.W. and Yan, H. Quantitative description of non-equilibrium turbulent phenomena in compressors, Aerospace Sci. Technol., 2017, 71, pp 7889.CrossRefGoogle Scholar
Wilcox, D.C. Turbulence Modeling for CFD, 3rd ed, DCW Industries, 2006, La Canada, CA.Google Scholar
Spalart, P.R. Philosophies and fallacies in turbulence modeling, Prog. Aerospace Sci., 2015, 74, pp 115.CrossRefGoogle Scholar
Park, J. and Choi, H. Toward neural-network-based large eddy simulation: Application to turbulent channel flow, J. Fluid Mech., 2021, 914, p A16.CrossRefGoogle Scholar
Kurz, M., Offenhäuser, P. and Beck, A. Deep reinforcement learning for turbulence modeling in large eddy simulations, Int. J. Heat Fluid Flow, 2023, 99, pp 109094.CrossRefGoogle Scholar
Pope, S.B. A more general effective-viscosity hypothesis, J. Fluid Mech., 1975, 72, (02), p 331.CrossRefGoogle Scholar
Ling, J., Kurzawski, A. and Templeton, J. Reynolds averaged turbulence modelling using deep neural networks with embedded invariance, J. Fluid Mech., 2016, 807, pp 155166.CrossRefGoogle Scholar
Zhang, X.L., Xiao, H., Luo, X.D. and He, G.W. Ensemble kalman method for learning turbulence models from indirect observation data, J. Fluid Mech., 2022, 949, p A26.CrossRefGoogle Scholar
Mandler, H. and Weigand, B. A realizable and scale-consistent data-driven non-linear eddy viscosity modeling framework for arbitrary regression algorithms, Int. J. Heat Fluid Flow, 2022, 97, p 109018.CrossRefGoogle Scholar
Weatheritt, J. and Sandberg, R. A novel evolutionary algorithm applied to algebraic modifications of the RANS stress–strain relationship, J. Comput. Phys., 2016, 325, pp 2237.CrossRefGoogle Scholar
Zhao, Y.M., Akolekar, H.D., Weatheritt, J., Michelassi, V. and Sandberg, R.D. RANS turbulence model development using CFD-driven machine learning, J. Comput. Phys., 2020, 411, p 109413.CrossRefGoogle Scholar
Lav, C., Banko, A.J., Waschkowski, F., Zhao, Y.M., Elkins, C.J., Eaton, J.K. and Sandberg, R.D. A coupled framework for symbolic turbulence models from deep-learning, Int. J. Heat Fluid Flow, 2023, 101, p 109140.CrossRefGoogle Scholar
Wu, J.L., Xiao, H. and Paterson, E. Physics-informed machine learning approach for augmenting turbulence models: A comprehensive framework, Phys. Rev. Fluids, 2018, 3, (7), p 074602.CrossRefGoogle Scholar
Brener, B.P., Cruz, M.A., Macedo, M.S.S. and Thompson, R.L. A highly accurate strategy for data-driven turbulence modeling, Comput. Appl. Math., 2024, 43, (1), p 59.CrossRefGoogle Scholar
Parish, E.J. and Duraisamy, K. A paradigm for data-driven predictive modeling using field inversion and machine learning, J. Comput. Phys., 2016, 305, pp 758774.CrossRefGoogle Scholar
Ferrero, A., Iollo, A. and Larocca, F. Field inversion for data-augmented rans modelling in turbomachinery flows, Comput. Fluids, 2020, 201, p 104474.CrossRefGoogle Scholar
Yan, C.Y., Li, H.R., Zhang, Y.F. and Chen, H.X. Data-driven turbulence modeling in separated flows considering physical mechanism analysis, Int. J. Heat Fluid Flow, 2022, 96, p 109004.CrossRefGoogle Scholar
Duraisamy, K., Zhang, Z.J. and Singh, A.P. New approaches in turbulence and transition modeling using data-driven techniques, 53rd AIAA Aerospace Sciences Meeting, 2015, p 1284.CrossRefGoogle Scholar
Yan, C.Y., Zhang, Y. and Chen, H. Data augmented turbulence modeling for three-dimensional separation flows, Phys. Fluids, 2022, 34, (7), p 075101.CrossRefGoogle Scholar
Zhu, L.Y., Zhang, W.W., Kou, J.Q. and Liu, Y.L. Machine learning methods for turbulence modeling in subsonic flows around airfoils, Phys. Fluids, 2019, 31, (1), p 015105.CrossRefGoogle Scholar
Zhang, S.M., Li, H.W., You, R.Q., Kong, T.L. and Tao, Z. A construction and training data correction method for deep learning turbulence model of Reynolds averaged Navier–Stokes equations, AIP Adv., 2022, 12, (6), p 065002.CrossRefGoogle Scholar
Wu, J.L., Xiao, H., Sun, R. and Wang, Q.Q. Reynolds-averaged navier–stokes equations with explicit data-driven reynolds stress closure can be ill-conditioned, J. Fluid Mech., 2019, 869, pp 553586.CrossRefGoogle Scholar
Liu, W.S., Fang, J., Rolfo, S., Moulinec, C. and Emerson, D.R. An iterative machine-learning framework for rans turbulence modeling, Int. J. Heat Fluid Flow, 2021, 90, p 108822.CrossRefGoogle Scholar
Brener, B.P., Cruz, M.A., Thompson, R.L. and Anjos, R.P. Conditioning and accurate solutions of reynolds average navier–stokes equations with data-driven turbulence closures, J. Fluid Mech., 2021, 915, p A110.CrossRefGoogle Scholar
Spencer, R., Przytarski, P., Adami, P., Grothe, P. and Wheeler, A.P.S. Importance of non-equilibrium modelling for compressors, J. Turbomach., 2022, pp 116.Google Scholar
Sun, X.X., Cao, W.B., Liu, Y.L., Zhu, L.Y. and Zhang, W.W. High Reynolds number airfoil turbulence modeling method based on machine learning technique, Comput. Fluids, 2022, 236, p 105298.CrossRefGoogle Scholar
Abe, H., Kawamura, H. and Matsuo, Y. Direct numerical simulation of a fully developed turbulent channel flow with respect to the Reynolds number dependence, J. Fluids Eng., 2001, 123, (2), pp 382393.CrossRefGoogle Scholar
Moser, R.D., Kim, J. and Mansour, N.N. Direct numerical simulation of turbulent channel flow up to ${\rm{R}}{{\rm{e}}_\tau }$ = 590, Phys. Fluids, 1999, 11, (4), pp 943945.CrossRefGoogle Scholar
Fang, L., Bao, T.W., Xu, W.Q., Zhou, Z.D., Du, J.L. and Jin, Y.. Data driven turbulence modeling in turbomachinery — an applicability study, Comput. Fluids, 2022, 238, p 105354.CrossRefGoogle Scholar
Duraisamy, K. Perspectives on machine learning-augmented reynolds-averaged and large eddy simulation models of turbulence, Phys. Rev. Fluids, 2021, 6, (5), p 050504.CrossRefGoogle Scholar
Schmelzer, M., Dwight, R.P. and Cinnella, P. Discovery of algebraic reynolds-stress models using sparse symbolic regression, Flow Turbul. Combust., 2020, 104, pp 579603.CrossRefGoogle Scholar
Akolekar, H.D., Weatheritt, J., Hutchins, N., Sandberg, R.D., Laskowski, G. and Michelassi, V. Development and use of machine-learnt algebraic reynolds stress models for enhanced prediction of wake mixing in low-pressure turbines, J. Turbomach., 2019, 141, (4).CrossRefGoogle Scholar
Mandler, H. and Weigand, B. On frozen-rans approaches in data-driven turbulence modeling: Practical relevance of turbulent scale consistency during closure inference and application, Int. J. Heat Fluid Flow, 2022, 97, p 109017.CrossRefGoogle Scholar
Parneix, S., Laurence, D. and Durbin, P.A. A procedure for using dns databases, J. Fluids Eng., 1998, 120, (1), pp 4047.CrossRefGoogle Scholar
Weatheritt, J. and Sandberg, R.D. The development of algebraic stress models using a novel evolutionary algorithm, Int. J. Heat Fluid Flow, 2017, 68, pp 298318.CrossRefGoogle Scholar
Wu, C.Y. and Zhang, Y.F. Enhancing the shear-stress-transport turbulence model with symbolic regression: A generalizable and interpretable data-driven approach, Phys. Rev. Fluids, 2023, 8, (8), p 084604.CrossRefGoogle Scholar
Liu, Y.W., Lu, L.P., Fang, L. and Gao, F. Modification of spalart–allmaras model with consideration of turbulence energy backscatter using velocity helicity, Phys. Lett. A, 2011, 375, (24), pp 23772381.CrossRefGoogle Scholar
Liu, Y.W., Tang, Y.M., Scillitoe, A.D. and Tucker, P.G. Modification of shear stress transport turbulence model using helicity for predicting corner separation flow in a linear compressor cascade, J. Turbomach., 2020, 142, (2), p 021004.CrossRefGoogle Scholar
Liu, Y.W., Yan, H., Fang, L., Lu, L.P., Li, Q.S. and Shao, L. Modified k- $\omega $ model using kinematic vorticity for corner separation in compressor cascades, Sci. China Technol. Sci., 2016, 59, (5), pp 795806.CrossRefGoogle Scholar
Sun, W. Assessment of advanced rans turbulence models for prediction of complex flows in compressors, Chin. J. Aeronaut., 2023, 36, (9), pp 162177.CrossRefGoogle Scholar
Klein, T.S., Craft, T.J. and Iacovides, H. Assessment of the performance of different classes of turbulence models in a wide range of non-equilibrium flows, Int. J. Heat Fluid Flow, 2015, 51, pp 229256.CrossRefGoogle Scholar
Liu, F., Lu, L.P. and Fang, L. Non-equilibrium turbulent phenomena in transitional channel flows, J. Turbul., 2018, 19, (9), pp 731753.CrossRefGoogle Scholar
Xiao, H., Wu, J.L., Laizet, S. and Duan, L. Flows over periodic hills of parameterized geometries: A dataset for data-driven turbulence modeling from direct simulations, Comput. Fluids, 2020, 200, p 104431.CrossRefGoogle Scholar
Wang, J.X., Wu, J.L. and Xiao, H. Physics-informed machine learning approach for reconstructing reynolds stress modeling discrepancies based on DNS data, Phys. Rev. Fluids, 2017, 2, (3), p 034603.CrossRefGoogle Scholar
Liu, W.S., Song, Z.M. and Fang, J. Nnpred: A predictor library to deploy neural networks in computational fluid dynamics software, arXiv preprint arXiv:2209.12339, September 2022.Google Scholar
Liu, W.S., Fang, J., Rolfo, S., Moulinec, C. and Emerson, D.R. On the improvement of the extrapolation capability of an iterative machine-learning based RANS framework, Comput. Fluids, 2023, 256, p 105864.CrossRefGoogle Scholar
Abadi, M., Barham, P., Chen, J.M., Chen, Z.F., Davis, A., Dean, J., Devin, M., Ghemawat, S., Irving, G., Isard, M., Kudlur, M., Levenberg, J., Monga, R., Moore, S., Murray, D.G., Steiner, B., Tucker, P., Vasudevan, V., Warden, P., Wicke, M., Yu, Y. and Zheng, X.Q. Tensorflow: A system for large-scale machine learning, Proceedings of the 12th USENIX Conference on Operating Systems Design and Implementation, OSDI’16, USA, 2016, USENIX Association, pp 265283. Google Scholar
Thompson, R.L., Sampaio, L.E.B., de Bragança Alves, F.A.V., Thais, L. and Mompean, G. A methodology to evaluate statistical errors in dns data of plane channel flows, Comput. Fluids, 2016, 130, pp 17.CrossRefGoogle Scholar
Poroseva, S.V., Colmenares, F.J.D. and Murman, S.M. On the accuracy of rans simulations with dns data, Phys. Fluids, 2016, 28, (11), p 115102.CrossRefGoogle ScholarPubMed
Vassilicos, J.C. Dissipation in turbulent flows, Ann. Rev. Fluid Mech., 2015, 47, (1), pp 95114.CrossRefGoogle Scholar
Bos, W.J.T. and Rubinstein, R. Dissipation in unsteady turbulence, Phys. Rev. Fluids, 2017, 2, (2), p 022601.CrossRefGoogle Scholar
Li, H.R., Zhang, Y.F. and Chen, H.X. Aerodynamic prediction of iced airfoils based on modified three-equation turbulence model, AIAA J., 2020, 58, (9), pp 38633876.CrossRefGoogle Scholar
Liu, F., Lu, L.P., Bos, W.J.T. and Fang, L. Assessing the nonequilibrium of decaying turbulence with reversed initial fields, Phys. Rev. Fluids, 2019, 4, (8), p 084603.CrossRefGoogle Scholar
Fang, L. and Bos, W.J.T. An EDQNM study of the dissipation rate in isotropic non-equilibrium turbulence, J. Turbul., 2023, 24, pp 217–324.CrossRefGoogle Scholar
Shao, X., Fang, J. and Fang, L. Non-equilibrium dissipation laws in a minimal two-scale wake model, Phys. Fluids, 2023, 35, p 085105.CrossRefGoogle Scholar
Araki, R. and Bos, W.J.T. Inertial range scaling of inhomogeneous turbulence, J. Fluid Mech., 2024, 978, p A9.CrossRefGoogle Scholar
Ma, B., Van Doorne, C.W.H., Zhang, Z. and Nieuwstadt, F.T.M. On the spatial evolution of a wall-imposed periodic disturbance in pipe poiseuille flow at $re$ = 3000. part 1. subcritical disturbance, J. Fluid Mech., 1999, 398, pp 181224.CrossRefGoogle Scholar
Xu, C., Zhang, Z., den Toonder, J.M.J. and Nieuwstadt, F.T.M. Origin of high kurtosis levels in the viscous sublayer. direct numerical simulation and experiment, Phys. Fluids, 1996, 8, (7), pp 19381944.CrossRefGoogle Scholar
Liu, F. Researches on the Non-Equilibrium Properties of Turbulence, PhD thesis, Beihang University, Beijing, China, 2019.Google Scholar
Hearst, R.J. and Lavoie, P. Velocity derivative skewness in fractal-generated, non-equilibrium grid turbulence, Phys. Fluids, 2015, 27, (7), p 071701.CrossRefGoogle Scholar
Hussain, A.K.M.F. and Reynolds, W.C. Measurements in fully developed turbulent channel flow, J. Fluids Eng., 1975, 97, (4), pp 568578.CrossRefGoogle Scholar
Pope, S.B. Turbulent Flows, Cambridge University Press, New York, 2000.Google Scholar
Zhao, H.K., Liu, Y.W., Shao, L., Fang, L. and Dong, M. Existence of positive skewness of velocity gradient in early transition, Phys. Rev. Fluids, 2021, 6, (10), p 104608.CrossRefGoogle Scholar
Tibshirani, R. Regression shrinkage and selection via the lasso, J. R. Stat. Soc. Ser. B (Methodolog.), 1996, 58, (1), pp 267288.CrossRefGoogle Scholar
Zhao, H.K. Analysis of Nonequilibrium Turbulence Phenomenon and Turbulence Modeling for Compressor Internal Model Flow, PhD thesis, Beihang University, Beijing, China, 2021.Google Scholar
Tao, S.C. and Zhou, Y. Turbulent flows around side-by-side cylinders with regular and multiscale arrangements, Phys. Rev. Fluids, 2019, 4, (12), p 124602.CrossRefGoogle Scholar
Figure 0

Figure 1. Comparisons of $\left( {{q_1},{q_2}} \right)$ in the transitional channel flow in transition region (TR), non-equilibrium region (NE) and equilibrium region (EQ) respectively [35].

Figure 1

Table 1. The input features and the target variable for the practical ML-RANS model [51, 52]

Figure 2

Figure 2. Mean velocity profiles by use of different high-fidelity methods.

Figure 3

Table 2. Information of LES on the spatially transitional channel flow [48]

Figure 4

Table 3. Hyperparameters of the neural network for the ML-RANS TR-NE-EQ model

Figure 5

Figure 3. Evolution of loss in the training process. The x-axis represents the ratio of current training step and total steps.

Figure 6

Figure 4. Profiles of the target variable $\nu _t^0$ at $x/( {{\pi}h} ) = 20,30,40,50$ in TR (blue), NE (red) and EQ (green) regions: comparison of ML predictions (lines) and a snapshot of training dataset (points).

Figure 7

Figure 5. Relation between the target variable $\nu _t^0$ and two input features $\left( {{q_1},{q_2}} \right)$ at four different streamwise locations: comparison of ML predictions (line) and a snapshot of training dataset (point).

Figure 8

Figure 6. Mapping from $\left( {{q_1},{q_2}} \right)$ to $\nu _t^0$ in two ML-RANS models.

Figure 9

Figure 7. Evolution of maximum value of the mean velocity along streamwise direction.

Figure 10

Figure 8. Sketch of cascade design of LPT: near-wall behaviours in simulations are observed both at the blue star (endwall, measurement along the $x$ axis) and at red arrows (blade).

Figure 11

Figure 9. Mesh of the channel with Blade 2 used in numerical simulations.

Figure 12

Figure 10. Comparison on results near endwall among the $k - \omega $ SST model and the two ML-RANS approaches, at the location marked as the blue star in Fig. 8: (a) Mean velocity, (b) Eddy viscosity.

Figure 13

Figure 11. Comparison on mean velocities near blade among the $k - \omega $ SST model and the two ML-RANS approaches, at the locations marked as the arrows with different distances from the leading edge in Fig. 8: (a) $0.2{l_{ax}}$, (b) $0.7{l_{ax}}$.

Figure 14

Figure 12. Comparison on $\mathrm{a}$-$\mathrm{posteriori}$ results on static pressure coefficient around the blade.

Figure 15

Figure 13. Mach number (Ma) calculated by (a) ML-RANS TR-NE-EQ model, (b) $k - \omega $ SST model and (c) ML-RANS EQ model at mid-span section: 0.3 at cyan contour and 0.7 at yellow contour.

Figure 16

Figure 14. Eddy viscosity (${\nu _t}$) calculated by (a) ML-RANS TR-NE-EQ model, (b) $k - \omega $ SST model and (c) ML-RANS EQ model at at mid-span section.

Figure 17

Figure 15. Mapping from $\left( {{q_1},{q_2}} \right)$ to $\nu _t^0$ of two ML-RANS models in the bar wake (${\nu _t} \gt 0.0015{{\rm{m}}^2}/{\rm{s}}$).

Figure 18

Figure 16. Total pressure loss coefficient at measurement plane on the blade mid-span section.

Figure 19

Figure 17. Mach number at measurement plane on the blade mid-span section.

Figure 20

Figure 18. Deviation of the second peak about the total pressure loss coefficient and the Mach number. The solid and larger symbols represent the result in the selected case (Blade 2 with zero angle-of-attack).