1 Introduction

When using computational fluid dynamics (CFD) to investigate complex, industrial relevant aircraft configurations at high Reynolds numbers, the direct simulation of turbulent fluctuations is still computationally prohibitive. Thus, typically, the Reynolds averaged Navier–Stokes (RANS) equations are solved and turbulent behavior is modelled using so called turbulence models. Various different models are available. However, all of them are based on coefficients which need to be tuned theoretically and empirically based on experimental data in order to obtain accurate predictions. Traditionally, this was and is a rather manual process in which one tries to match numerical results with canonical experiments without accounting for uncertainties. Due to increasing computational resources and data being available, data-driven approaches based on machine learning gained attention for such coefficient tuning tasks within the past few years [1]. The idea of these data-driven approaches is to re-infer and/or to adjust the established coefficients or terms in the turbulence model based on experimental data or numerical data computed by direct numerical simulations (DNS) or other higher-fidelity numerical techniques by using statistical methods.

In CFD computations, the issue of physical modeling inaccuracy is normally inevitable. Numerical errors due to spatial and temporal discretization of the governing equations and physical models are one of the most important issues in CFD computations [2]. These inaccuracy and errors are recognized as epistemic uncertainty, which can be reduced if more knowledge is gained. Turbulence models are one of the important issues here. In measurements, on the other hand, the observed data may contain measurement errors due to various unknown external factors. These unknown factors are classified as aleatory uncertainty, which is inherently regarded as irreducible.

Data-driven methods depend on our hypothesis in statistical inference. One example of a conceivable hypothesis is that the form of the turbulence closure model is accurate enough, but that the coefficients defined in empiricism that relate to specific flow properties are not always correct for a variety of flow cases due to insufficient knowledge. In this article these coefficients are treated as stochastic by Bayesian inference. Bayesian inference refers to statistical inference where uncertainty in inferences is quantified using probability. In general, statistical inference methods vary from problem to problem but can abstractly be grouped in two categories on how to treat the parameters (latent variables), which are the coefficients in closure models. On the one hand, if large data sets are available, so-called point-estimate approaches are commonly applied with deep learning being the arguably most promising and efficient methodology. On the other hand, the inference of the latent variables by regarding them as stochastic by enforcing a Bayesian perspective provides uncertainty quantification (UQ) capabilities associated with the latent variables and enables another perspective on the quantities of interest (QoIs) in terms of observable variables. These capabilities become especially remarkable when the amount of data is relatively small. However, effective numerical techniques for UQ are crucial especially in high dimensional cases. In this article, the closure coefficients are treated as stochastic latent variables. The expression latent variables refers to unobserved variables and is used as an antonym to observed variables such as input and output variables. The stochastic approach then enables to quantify uncertainties of the latent variables in the QoI in terms of credible intervals. The above-mentioned categories are rigorously classified in Sect. 2. In this article, to emphasize the differences between tuning parameters by a point-estimate and evaluating them stochastically by an interval-estimate, we sometimes use the expressions calibration and inference, respectively. Here we start with reviewing some previous work on the point-estimate approach.

One of the examples of turbulence model calibration with the point-estimate approach is to minimize the error between experimental data and the output of RANS-based CFD computations for some QoIs. This is a direct approach. From a statistical inference perspective, this is known as a parametric regression process or parameter calibration: the turbulence model is considered as a nonlinear regression model whose inputs and outputs are parameters such as flow conditions, the numerical mesh, etc., and QoIs such as integral aerodynamic coefficients or pressure distributions, respectively. The closure coefficients are treated as the parameters (or globally speaking, latent variables) of the nonlinear regression model. Thus, least-squares optimization through minimizing the mean squared error (MSE) [3] provides turbulence closure coefficients calibration. In general, during the calibration process a larger number of RANS simulations are required causing a huge computational burden especially for industrial relevant cases. Hence, large computational costs during the optimization can be partly circumvented by relying on surrogate models such as Gaussian processes or neural networks [3, 4]. Another alternative is the use of adjoint flow solvers to efficiently compute the gradients of the QoIs with respect to the closure coefficients, enabling efficient gradient-based optimization for solving the least-square problems.

Another example of approaches using the point-estimate is calibration of the turbulence closure coefficients as the output of the statistical inference methods [5,6,7]. Since the coefficients to be inferred are now the output of regression models, not parameters in the models, any kind of supervised learning techniques are applicable. For example, the authors in Refs. [5,6,7] used neural networks to handle millions of data points by considering spatio-temporal parameters on flow fields.

These two approaches introduced above can theoretically be extended to an interval-estimate via the Bayesian perspective to evaluate the uncertainty associated with the closure coefficients. For example, in the direct approach using the least squares, the closure coefficients as the latent variables can be treated as stochastic to formulate a posterior distribution of the coefficients. As a result, the output QoIs also form a probability distribution, which is called predictive distribution. These probability distributions cannot be obtained analytically because the probabilistic models behind are inherently nonlinear regression models. The second approach mentioned in above may have further difficulties in extension to the interval-estimate. In case the parameters to be calibrated are high-dimensional, the use of Gaussian processes may become infeasible [7]. Bayesian neural networks would have to rely on approximation methods to compute the posterior distribution because of the high dimensionality of the parameters. In this article, our approach is based on the direct approach that the nonlinear regression model is a turbulence model equation with closure coefficients as parameters (the latent variables) to be inferred using the interval-estimate via the Bayesian perspective.

There are several methods to compute the posterior and the predictive distributions. The variational inference and the Laplace approximation [8] are versatile methods to approximate the posterior distribution by a parametric probability distribution. Since both of them require an iterative optimization process with many RANS computations, the direct simulation is still not as practical as the calibration process by the least squares. Markov-chain Monte Carlo (MCMC) methods [8, 9] are used to approximate the posterior distribution of a parameter of interest by random sampling in a parameter space. Unlike the above-mentioned parametric approximation methods, MCMC provides a theoretical guarantee that the histograms of the obtained sample points compose the posterior distribution when the sample size approaches infinity. Since each sample requires a RANS computation and the number of samples is usually greater than the number of iterations of the optimization, a direct implementation of MCMC is infeasible. It is essential for all MCMC methods to be supported by surrogate models in order to reduce the number of RANS computations. Once a surrogate model is constructed, the computational cost of MCMC can be negligible and MCMC can be regarded to have higher fidelity for computing the posterior than the other methods. However, it has to be noted that the application of MCMC to high-dimensional spaces remains a challenging topic. Also, from the point of view of surrogate model construction, it is desirable to keep the dimension of the parameters to be inferred as small as possible.

There is a body of existing research on surrogate-assisted MCMC approaches [10,11,12]. In Ref. [12], output uncertainty is directly computed via the samples generated by MCMC using an efficient boundary layer code. In Ref. [11] MCMC is used to evaluate the probability in the Bayesian update process for the purpose of calibrating turbulence model parameters. Surrogate-assisted MCMC approaches are also used for sensitivity analysis in the context of UQ [3, 13]. A review of recent work related to turbulence models can be found in Ref. [14].

In this article, the uncertainty associated with the closure coefficients is rigorously evaluated with a help of MCMC. The quantified uncertainty in the QoIs can be identified intuitively as a lack of knowledge of the coefficients. On the other hand, there are other uncertainties lurking in flow physics as mentioned above. All the other uncertainties, no matter where they come from, are classified as noise in a Gaussian probabilistic model to focus on the uncertainty associated with the closure coefficients.

The probabilistic model is a parametric model governed by a model parameter that describes the set of closure coefficients and a parameter to control the noise. The main feature of the Bayesian perspective here is to regard the parameters (the latent variables) as stochastic variables. In order to fully execute the Bayesian approach to account for all the uncertainties that are present, all the parameters are treated as stochastic. This modeling approach can quantify the uncertainty in the QoIs that stems from the contribution of all the uncertainties, i.e., uncertainty due to the closure coefficients and due to noise. However, our interest is to quantify the uncertainties associated with the closure coefficients. From a physical perspective, this uncertainty needs to be clearly distinguished from the noise. However, the parameter to control the noise depends on the turbulence model parameters when data is observed (it forms a joint probability). A method for extracting the uncertainty associated with the closure coefficients is presented and demonstrated using various test cases. Note that modeling only the closure coefficients as stochastic variables yields different results which do not account for the effects of the uncertainty due to noise.

As mentioned earlier, the ability to quantify the epistemic uncertainty associated with closure coefficients has a number of potential applications, for example, feedback to experimental measurement techniques and propagation of the uncertainty to the output QoI for arbitrary flow types. It can be possible to identify at which domains in the input parameter space (such as angles of attack, Mach number, and the location on the wings) we need to measure the output QoI in experiments (sensor placement). Note that this process is basically the same concept as so-called Bayesian optimization and active learning, effectively employing the posterior uncertainty. There is, however, an important difference as the inferred parameters (the latent variables) are physically meaningful in our applications.

The target test case to demonstrate the data-driven Bayesian inference of the turbulence model closure coefficients is based on the Airbus XRF1 research aircraft configuration and corresponding wind tunnel data at flight Reynolds numbers. The RAE2822 airfoil is also used to verify the methodology, but with numerically generated measured data. The output QoIs are the drag and surface pressure coefficients, for which measured data is required to infer the model coefficients.

The flow solver and the turbulence model used in this article, the methodologies and the developed framework are introduced in Sect. 2. In Sect. 3, results are presented for two test cases and discussed. Section 4 provides a summary and conclusions.

2 Methodology

This section introduces a framework for data-driven turbulence modeling and provides a generalized perspective to showcase that the developed framework can be applied to other (turbulence) models as well as other test cases.

2.1 Flow solver and turbulence model

The flow solver employed is the DLR TAU-code. It is based on a finite volume scheme, where inviscid terms are computed via a first or second order scheme. Viscous terms are computed via a second order scheme. Details of the solver can be found in Ref. [15,16,17]. In this paper, matrix dissipation is used. For time integration a LUSGS implicit scheme is chosen. Convergence acceleration is achieved with a multi-grid algorithm based on agglomerated coarse grids. Various turbulence models are available and the negative Spalart–Allmaras turbulence model [18] (referred to as the negative S–A) is investigated and hence altered in this paper.

The equations of the negative S–A are briefly described next. The negative S–A is derived from the original Spalart–Allmaras turbulence model (referred to as the original S–A). Details of the derivations provided below are available in Ref. [18]. The original S–A is a linear eddy viscosity model based on Boussinesq approximation. The turbulent eddy viscosity \({\nu}_{t}\) is calculated by:

$${\nu_{t} = \tilde{\nu }f_{v1} .}$$
(1)

Here, the transport equation to be solved to obtain \(\tilde{\nu }\) is as:

$$\frac{{{\text{D}}\tilde{\nu }}}{{{\text{D}}t}} = \underbrace {{c_{b1} \left( {1 - f_{t2} } \right)\tilde{S}\tilde{\nu }}}_{{{\text{production}}}} - \underbrace {{\left( {c_{w1} f_{w} - \frac{{c_{b1} }}{{\kappa^{2} }}f_{t2} } \right)\left( {\frac{{\tilde{\nu }}}{d}} \right)^{2} }}_{{{\text{destruction}}}} + \cdots \underbrace {{\frac{1}{\sigma }\left[ {\nabla \cdot \left( {\left( {\nu + \tilde{\nu }} \right)\nabla \tilde{\nu }} \right) + c_{b2} \left( {\nabla \tilde{\nu }} \right)^{2} } \right],}}_{{{\text{diffusion}}}}$$
(2)

where \(\tilde{S}\) is the modified vorticity,

$${\tilde{S} = S + \overline{S},\quad \overline{S} = \frac{\nu }{{\kappa^{2} d^{2} }}\chi f_{v2} ,}$$
(3)

where \(S\) is the vorticity magnitude. The other coefficients in Eqs. (13) are defined as:

$${\left\{ {\begin{array}{*{20}l} {c_{w1} = \frac{{c_{b1} }}{{\kappa^{2} }} + \frac{{1 + c_{b2} }}{\sigma },} \\ {f_{v1} = \frac{{\chi^{3} }}{{\chi^{3} + c_{v1}^{3} }},} \\ {\chi = \frac{{\tilde{\nu }}}{\nu },} \\ {f_{v2} = 1 - \frac{\chi }{{1 + \chi f_{v1} }},} \\ {f_{w} = g\left( {\frac{{1 + c_{w3}^{6} }}{{g^{6} + c_{w3}^{6} }}} \right)^{1/6} ,} \\ {g = r + c_{w2} \left( {r^{6} - r} \right),} \\ {r = \mathop {\min }\limits_{{}} \left( {\frac{{\tilde{\nu }}}{{\tilde{S}\kappa^{2} d^{2} }},10} \right),} \\ {f_{v1} = c_{t3} exp\left( { - c_{t4} \chi^{2} } \right).} \\ \end{array} } \right.}$$
(4)

The original S–A contains two additional coefficients namely \(c_{t1}\) and \(c_{t2}\). In the DLR TAU-code, the trip term in Eq. (2) is neglected. Therefore, 9 closure coefficients are accounted for.

In the negative S–A, there are two modifications to the original S–A. The first one is that the modified vorticity \(\tilde{S}\) in Eq. (3) is enforced to be positive. The other one is a modification of the diffusion coefficient \(\nu + \tilde{\nu }\) in Eq. (2). Therefore \(\overline{S}\) in Eq. (2) is represented by two additional coefficients \(c_{v2}\) and \(c_{v3}\) when \(\overline{S} \leftarrow c_{v2} S\) as:

$${\tilde{S} = S + \frac{{S\left( {c_{v2}^{2} S + c_{v3} \overline{S}} \right)}}{{\left( {c_{v3} - 2c_{v2} } \right)S - \overline{S}}}.}$$
(5)

Otherwise Eq. (3) is retained. The diffusion coefficient \(\nu + \tilde{\nu }\) in Eq. (2) is modified to be \(\nu + \tilde{\nu }f_{n}\), where \(f_{n}\) is expressed as follows by introducing one coefficient \(c_{n1}\):

$${f_{n} = \frac{{c_{n1} + \chi^{3} }}{{c_{n1} - \chi^{3} }}.}$$
(6)

Hence, the negative S–A includes 12 closure coefficients. These coefficients are regarded as a parameter vector which should be inferred. Moreover, constraints on the feasible intervals of the coefficients are introduced to ± 50% of the default values. Table 1 shows the original coefficient values of the negative S–A and the valid intervals for the inference processes. Note that in Ref. [18], the modified vorticity \(\tilde{S}\) in the negative S–A should not below 0.3\(S\) posing an additional implicit constraint. Equations (3) and (5), might introduce further constraints on the intervals on \(c_{v2}\) and \(\kappa\), but these restrictions are not considered.

Table 1 Default values and intervals of the negative Spalart–Allmaras turbulence model for the inference processes

2.2 Theory

This sub-section introduces a parametric statistical inference method as the cornerstone of the developed framework. The process of the statistical inference method is fundamentally composed of two steps, learning and prediction. These steps are conducted on a parametric probabilistic model that we define. The definition relies on two hypotheses as the first two steps. Eventually the process of the parametric statistical inference is in a broad way composed of the following four steps:

  1. 1.

    Define a regression model

  2. 2.

    Define a probabilistic model

  3. 3.

    Compute a posterior by given data (learning process)

  4. 4.

    Compute output distributions (prediction process)

The first two steps introduce the two hypotheses by each. The overall process here can be also regarded as extension of Bayesian linear regression to nonlinear regression models in general. A principal figure to illustrate the process is shown Fig. 1. The goal is to compute the posterior distribution \(p\left( {{\varvec{a}},\sigma {|}{\mathbf{X}},{\mathbf{Y}}} \right)\) and the predictive distribution contributed only by the closure coefficients expressed as \(\left. {p\left( {y{|}{\varvec{x}}, {\mathbf{X}},{\mathbf{Y}}} \right)} \right|_{{\varvec{a}}}\) in Fig. 1. Details of each step are described in the next sub-sections.

Fig. 1
figure 1

Principal figure of the statistical inference process

2.2.1 The 1st Step: defining a regression model

The regression model here means a deterministic function that produces a scalar output \(y\) when an arbitrary input vector \({\varvec{x}}\) is given. The model characteristics are controlled by a parameter vector \({\varvec{a}}\). These are deterministic variables currently. Normally the parameter \({\varvec{a}}\) is being determined by the learning process in the 3rd step with available data. Based on this concept, a CFD computation using a turbulence model can be represented as the deterministic function device:

$${y = f\left( {{\varvec{x}},{\varvec{a}}} \right),}$$
(7)

where \({\varvec{a}}\) represents a vector as a set of all the turbulence model coefficients. The input \({\varvec{x}}\) is a vector as a set of flow conditions and/or a spatial coordinate. The output \(y\) is a scalar value as an output aerodynamic value such as lift and drag or distributed quantities, e.g. pressure coefficient distributions. The function \(f\) represents then the CFD computation with a turbulence model as a mapping \(f:\left( {{\varvec{x}},{\varvec{a}}} \right) \mapsto y\). For example, in this paper a turbulence model here is the negative S–A. This is the first hypothesis in the statistical inference method. As this first hypothesis for example, other turbulence models could be set here by modeling it as \(f\left( {{\varvec{x}},{\varvec{a}}} \right)\). The negative S–A has 12 independent coefficients, which are summarized as one parameter vector \({\varvec{a}}\), whose dimensionality is 12, i.e. \({\varvec{a}} = \left( {\sigma_{SA} ,\kappa ,c_{b1} ,c_{b2} ,c_{v1} ,c_{v2} ,c_{v3} ,c_{w2} ,c_{w3} ,c_{ct3} ,c_{ct4} ,c_{n1} } \right)\) by following Table 1. The parameter \({\varvec{a}}\) in Eq. (7) is called model parameter or simply parameter in this paper. Note that when our interest is only the input and the output as supervised learning, the function \(f\) is freely chosen (e.g. by model selection conducted in the third and fourth steps).

2.2.2 The 2nd Step: defining a probabilistic model

In the second step, now the output \(y\) is considered to be a stochastic variable. The process starts from assuming that the observed data \(y\) as the output has been generated by a probabilistic model as \(y\sim p\left( y \right)\). Our interest is evaluating the epistemic uncertainty associated with the turbulence model \(y\sim p\left( y \right)\) by diminishing the effects of all the other uncertainties, even the mesh and the scheme effects. Other possible strategies for Bayesian inference can be found in Ref. [19]. First, the discrepancy between the output \(y\) and the model \(f\left( {{\varvec{x}},{\varvec{a}}} \right)\) is represented by introducing a stochastic variable labelled \(\varepsilon\). Thus, the output \(y\) can be defined as follows:

$${y: = \underbrace {{f\left( {{\varvec{x}},{\varvec{a}}} \right)}}_{{\begin{array}{*{20}l} {the\,\,source\,\,term\,\,of} \\ {the\,\,uncertainty} \\ {associated\,\,with\,\,} {\varvec{a}} \\ \end{array} }} + \underbrace {\varepsilon }_{{\begin{array}{*{20}l} {the\,source\,term\,of} \\ {all\,the\,other} \\ {uncertainties} \\ \end{array} }}.}$$
(8)

The physical meaning of the stochastic variable \(\varepsilon\) is noise in the output value \(y\). Since \(\varepsilon\) is stochastic, a probabilistic model is introduced into this variable. One example of the probabilistic model of \(\varepsilon\) is an univariate Gaussian distribution as \(\varepsilon \sim {\mathcal{N}}\left( {0,\sigma^{2} } \right)\), where \({\mathcal{N}}\) and \(\sigma^{2}\) represents the Gaussian distribution and its variance, respectively. This eventually brings the following conditional probability distribution of \(y\):

$${p\left( {y{|}{\varvec{x}},{\varvec{a}},\sigma } \right) = {\mathcal{N}}\left( {y{|}f\left( {{\varvec{x}},{\varvec{a}}} \right),\sigma^{2} } \right).}$$
(9)

Now \(\sigma\) has been added as another parameter (called hyperparameter or noise parameter in this paper to be distinguished with the parameter \({\varvec{a}}\)). Equation (9) is the probabilistic model of the observed variable \(y\) as the objective here. Different from the observed variable \(y\) (and \({\varvec{x}}\)), the parameters \({\varvec{a}}\) and \(\sigma\) are not observed (the latent variables). In the fully Bayesian approach, the hyperparameter \(\sigma\) is also treated as a stochastic variable as \(\sigma \sim p\left( \sigma \right)\). Thus, the defined probabilistic model can account for all the uncertainties that are present.

Following Eq. (8), two sorts of induced uncertainties can be easily separated by looking at each term individually. since the model parameter \({\varvec{a}}\) is the only parameter that can directly interact with the internal structures of the model \(f\). All the other external uncertainties are dealt within the term \(\varepsilon\) governed by the noise parameter \(\sigma\). No other functions (such as bias functions) are added in Eq. (8). The parameter \({\varvec{a}}\) is then treated as a stochastic variable in the Bayesian approaches as \({\varvec{a}}\sim p\left( {\varvec{a}} \right)\). The function \(f\) as the first term also becomes stochastic. The stochastic variable \(y\sim p\left( y \right)\) as the predictive distribution becomes a mixture of probabilities derived from the two stochastic terms after the learning process. After the learning process when data is given, all the parameters in Eq. (9) as the latent variables are not independent of each other and compose a joint probability distribution. In the 4th step, a rigorous form of the probability of \(y\) as \(p\left( y \right)\) is presented. In that form, the contribution of the uncertainty associated with the parameter \({\varvec{a}}\) is not that obvious like Eq. (8) due to the above fact.

2.2.3 The 3rd Step: computing a posterior by given data (learning process)

The third step is the learning process of the parameters \({\varvec{a}}\) and \(\sigma\) using data. In this step, derivation to the Bayesian approach is employed. The process in this step is to evaluate the joint probability of all the parameters \({\varvec{a}}\) and \(\sigma\) as a posterior distribution. The posterior distribution can be evaluated using the Bayes’ theorem by prior distributions of all the parameters \({\varvec{a}}\) and \(\sigma\), and the likelihood function obtained by the probabilistic model and the data. How to compute the posterior function is described below. Some additional details about the definition of the prior distributions etc. are in Appendix 1.

The likelihood function can be defined by data and the probabilistic model defined in the second step as:

$${p\left( {{\mathbf{Y}}{|}{\mathbf{X}},{\varvec{a}},\sigma } \right) = \mathop \prod \limits_{i = 1}^{N} {\mathcal{N}}\left( {y^{\left( i \right)} {|}f\left( {{\varvec{x}}^{\left( i \right)} ,{\varvec{a}}} \right),\sigma^{2} } \right),}$$
(10)

where \({\mathbf{X}} = \left( {{\varvec{x}}^{\left( 1 \right)} ,{\varvec{x}}^{\left( 2 \right)} , \ldots ,{\varvec{x}}^{\left( N \right)} } \right)\) and \({\mathbf{Y}} = \left( {y^{\left( 1 \right)} ,y^{\left( 2 \right)} , \ldots ,y^{\left( N \right)} } \right)\) are observed data. \(N\) denotes the sample size of the data. The data is assumed to be independently drawn from the defined probabilistic model by Eq. (9) since the model is an univariate Gaussian distribution, hence no correlations exist. It has to be emphasized that direct evaluation of the likelihood function by Eq. (10) is practically available for the learning process instead of that of the posterior \(p\left( {{\varvec{a}},\sigma {|}{\mathbf{X}},{\mathbf{Y}}} \right)\) (see Appendix 1). The computation of the marginal likelihood \(p\left( {{\mathbf{Y}}{|}{\mathbf{X}}} \right)\) used as the normalization constant is either not necessary in executing the MCMC sampling techniques.

There are mainly two perspectives in the learning process (also see Fig. 1). One is point-estimate while the other is interval-estimate. In practice, these names are driven from the properties of the parameters to be inferred using the posterior distribution. The commonly known least square (LSQ) approach can be categorize as a special case of the point-estimate. The interval-estimate simply corresponds to using all the information of the posterior itself. The point-estimate in this paper directly indicates maximum a posterior (MAP). The MAP can be achieved by literally maximizing a posterior function \(p\left( {{\varvec{a}},\sigma {|}{\mathbf{X}},{\mathbf{Y}}} \right)\) with respect to the parameters \({\varvec{a}}\) and \(\sigma\):

$${\hat{\user2{a}},\hat{\sigma } = \mathop {{\text{argmax}}}\limits_{{{\varvec{a}},\sigma }} p\left( {{\varvec{a}},\sigma {|}{\mathbf{X}},{\mathbf{Y}}} \right).}$$
(11)

In practice, Eq. (11) is solved analytically when the regression model \(f\left( {{\varvec{x}},{\varvec{a}}} \right)\) is a linear regression model. Since the regression model here is a nonlinear regression model, typically Eq. (11) is solved employing optimizers.

Under the above-mentioned condition that no specific prior distribution is set, the direct evaluation of the likelihood function by Eq. (10) is available in practice. When pursuing the point-estimate using an optimizer, error functions which are derived by taking the negative of the log-likelihood function (see Appendix 5 for further details) are practically used. When pursuing the interval-estimate, the sampling method using MCMC can be directly applied to the likelihood function in Eq. (10). This evaluation by MCMC provides an approximation of the posterior \(p\left( {{\varvec{a}},\sigma {|}{\mathbf{X}},{\mathbf{Y}}} \right)\) by sample points. The posterior is composed by a histogram of the sample points in the parameters space \({\varvec{a}},\sigma\). This means that the frequency of the generated sample points directly forms a scaled function of the posterior. The mode of the posterior \(p\left( {{\varvec{a}},\sigma {|}{\mathbf{X}},{\mathbf{Y}}} \right)\) represents a pair of the calibrated parameters \(\hat{\user2{a}},\hat{\sigma }\). By using the obtained sample points, the predictive distributions \(\left. {p\left( {y{|}{\varvec{x}}, {\mathbf{X}},{\mathbf{Y}}} \right)} \right|_{{\varvec{a}}}\) shown in Fig. 1 as our goal are computed at the fourth step. Again note that same as requirement of optimizers in Eq. (11), the reason why numerical approaches are required is because the regression model \(f\left( {{\varvec{x}},{\varvec{a}}} \right)\) is a nonlinear regression model.

Various established algorithms for MCMC are available and herein the random walk Metropolis-Hastings is applied. Details on MCMC algorithms can be found in many papers [8, 9]. As for optimizer for the point-estimate, we use differential evolution algorithm, which is a heuristic method. The reason why we do not use only the MCMC even for the point-estimate is for double-check of the accuracy of the tools. Therefore, the results by optimizers are not directly used in the prediction process.

2.2.4 The 4th Step: computing output distributions (prediction process)

In the last step the output \(y\) is computed by using the learned parameters: \(\hat{\user2{a}},\hat{\sigma }\) by the point-estimate, or \(p\left( {{\varvec{a}},\sigma {|}{\mathbf{X}},{\mathbf{Y}}} \right)\) by the interval-estimate. First as for the point-estimate, the prediction of the output \(y\) for a given input \({\varvec{x}}\) is obtained by evaluating the regression model in Eq. (7):

$${y = f\left( {{\varvec{x}},\hat{\user2{a}}} \right).}$$
(12)

Note that even for the point-estimate, the predictive distribution \(p\left( {y{|}{\varvec{x}}, {\mathbf{X}},{\mathbf{Y}}} \right)\) is introduced following Eq. (9). However, this predictive distribution contains only the noise \(\varepsilon\) as identified by the point-estimate represented by the standard deviation \(\hat{\sigma }\) as \(\varepsilon \sim {\mathcal{N}}\left( {0,\hat{\sigma }^{2} } \right)\).

In the interval-estimate (with fully Bayesian perspective), the parameters \({\varvec{a}},\sigma\) form a joint probability as \(p\left( {{\varvec{a}},\sigma {|}{\mathbf{X}},{\mathbf{Y}}} \right)\) which is represented by the MCMC sample points. The predictive distribution \(p\left( {y{|}{\varvec{x}}, {\mathbf{X}},{\mathbf{Y}}} \right)\) is obtained as follows by the sum and product rules of the probability theory:

$$\begin{aligned} p\left( {\left. y \right|\user2{x},~{\mathbf{X}},{\mathbf{Y}}} \right) & = \int{\!\!\int {p\left( {\left. y \right|\user2{x},\user2{a},\sigma } \right)p\left( {\user2{a},\left. \sigma \right|{\mathbf{X}},{\mathbf{Y}}} \right){\text{d}}\user2{a}{\text{d}}\sigma } } \\ & \approx \frac{1}{{N_{{mcmc}} }}\mathop \sum \limits_{{i = 1}}^{{N_{{{{mcmc}}}} }} p\left( {\left. y \right|\user2{x},\left( {\user2{a},\sigma } \right)^{{\left( i \right)}} } \right) \\ \end{aligned} $$
(13)

The parameters \({\varvec{a}},\sigma\) are marginalized (integrated out) to be eliminated. This operation can be regarded as taking the expectation. The probability as the weight is already realized by the MCMC sampling result. Therefore, it can be computed by mixtures of the probabilistic model \(p\left( {y{|}{\varvec{x}},{\varvec{a}}^{\left( i \right)} ,\sigma^{\left( i \right)} } \right)\), where each sample point \(i\) is weighted with its respective probability \(p\left( {{\varvec{a}}^{\left( i \right)} ,\sigma^{\left( i \right)} {|}{\mathbf{X}},{\mathbf{Y}}} \right)\). The constraint that the sum of all the weights has to be 1 to be a probability is accomplished by being divided with the sample size \(N_{mcmc}\). Note that, the predictive distribution shown in Eq. (13) is a common way for a fully Bayesian approach [8]. No hyperparameters in priors of the parameters \({\varvec{a}}\) and \(\sigma\) are explicitly presented in Eq. (13) since an improper prior for both \({\varvec{a}}\) and \(\sigma\) is assumed. Note that investigation when using an empirical Bayes approach in contrast to the full Bayes approach is shown in Appendices 2 and 3. In Appendix 3, a comparison of results between \(p\left( {y{|}{\varvec{x}}, {\mathbf{X}},{\mathbf{Y}}} \right)\) in Eq. (13) and \(\left. {p\left( {y{|}{\varvec{x}}, {\mathbf{X}},{\mathbf{Y}}} \right)} \right|_{{\varvec{a}}}\) in Eq. (15a, 15b) using the empirical Bayes approach is also shown.

The predictive distribution \(p\left( {y{|}{\varvec{x}}, {\mathbf{X}},{\mathbf{Y}}} \right)\) in Eq. (13) represents the contribution of all the sources of the uncertainties, one of which is associated with the model parameter \({\varvec{a}}\), and all the others of which is associated with the noise parameter \(\sigma\). These two sorts of uncertainties physically stem from different sources as shown in Eq. (8) and the predictive distribution \(p\left( {y{|}{\varvec{x}}, {\mathbf{X}},{\mathbf{Y}}} \right)\) in Eq. (13) lost the information of their origins. Our goal is to quantify the epistemic uncertainty associated with the closure coefficients \({\varvec{a}}\), without including the external factors represented by the noise \(\varepsilon \sim {\mathcal{N}}\left( {0,\sigma^{2} } \right)\). This is accomplished through the following two steps by using the sum and product rules of the probability theory. The first step is to obtain the marginal posterior \(p\left( {{\varvec{a}}{|}{\mathbf{X}},{\mathbf{Y}}} \right)\) by marginalizing the posterior \(p\left( {{\varvec{a}},\sigma {|}{\mathbf{X}},{\mathbf{Y}}} \right)\) with respect to \(\sigma\):

$${p\left( {{\varvec{a}}{|}{\mathbf{X}},{\mathbf{Y}}} \right) = \int p\left( {{\varvec{a}},\sigma {|}{\mathbf{X}},{\mathbf{Y}}} \right){\text{d}}\sigma .}$$
(14)

Then, in order not to take the noise term in Eq. (8) into account, the probabilistic model of Eq. (9) where \(\sigma \to 0\) is considered. Eventually the predictive distribution where only the contribution of the parameter \({\varvec{a}}\) is considered is represented as follows:

$$\begin{aligned} \left. {p\left( {y{|}{\varvec{x}}, {\mathbf{X}},{\mathbf{Y}}} \right)} \right|_{{\varvec{a}}} &= \int p\left( {y{|}{\varvec{x}},{\varvec{a}}} \right)p\left( {{\varvec{a}}{|}{\mathbf{X}},{\mathbf{Y}}} \right){\text{d}}{\varvec{a}} \\ &\approx \frac{1}{{N_{mcmc} }}\mathop \sum \limits_{i = 1}^{{N_{mcmc} }} p\left( {y{|}{\varvec{x}},{\varvec{a}}^{\left( i \right)} } \right)\end{aligned}$$
(15a)
$${\text{with}}\;p\left( {y{|}{\varvec{x}},{\varvec{a}}} \right) = \delta \left( {y - f\left( {{\varvec{x}},{\varvec{a}}} \right)} \right),$$

where \(p\left( {{\varvec{a}}{|}{\mathbf{X}},{\mathbf{Y}}} \right)\) indicates the marginal distribution obtained by Eq. (14), \(\delta\) is the Dirac delta function as the limit of normal distributions as \(\sigma \to 0\). In practice, the technical process itself becomes eventually very simple. The predictive distribution \(\left. {p\left( {y{|}{\varvec{x}}, {\mathbf{X}},{\mathbf{Y}}} \right)} \right|_{{\varvec{a}}}\) in Eq. (15a) can be approximated by the output values of the MCMC sample points themselves in practice (by substituting the MCMC samples for the regression model \(f\left( {{\varvec{x}},{\varvec{a}}} \right)\)):

$${f:\left( {{\varvec{x}},{\varvec{a}}^{\left( i \right)} } \right) \mapsto y^{\left( i \right)} \left( {i = 1,2, \cdots ,N_{mcmc} } \right),}$$
(15b)

where \({\varvec{a}}_{i}\) is the \(i\)th sample point \(\left( {{\varvec{a}},\sigma } \right)_{i}\) obtained by the MCMC sampling with omitting the \(\sigma_{i}\) component in the sample point. The meaning of the predictive distribution \(\left. {p\left( {y{|}{\varvec{x}}, {\mathbf{X}},{\mathbf{Y}}} \right)} \right|_{{\varvec{a}}}\) in Eq. (15a) is different from that of \(p\left( {y{|}{\varvec{x}}, {\mathbf{X}},{\mathbf{Y}}} \right)\) in Eq. (13). Both of them are a probability of the output \(y\) after learning from the same data. However, the predictive distribution in Eq. (15a) represents the output uncertainty contributed only by the turbulence closure coefficients \({\varvec{a}}\) which have been learned by considering all the uncertainties that are present. The closure coefficients \({\varvec{a}}\) themselves were learned by taking the information of the uncertainty of the noise into account but the output \(y\) as QoI reflects only the epistemic uncertainty associated by \({\varvec{a}}\) eventually.

Overall for the entire steps, the process is common when a likelihood function is determined. The likelihood function in Eq. (10) is one example. Based on the concept of the four steps, the likelihood function to be evaluated (by either an optimizer or MCMC) can be naturally and flexibly determined for each problem setting. Note that further abstraction of the entire process of the four steps can be generalized by classifying all the stochastic variables into the latent variables and the observed variables (see Appendix 6).

2.3 Practical implementation details

This section provides insight into the practical implementation details with a special focus on computing the likelihood function. The likelihood function represented by Eq. (10) contains \(f\left( {{\varvec{x}}^{\left( i \right)} ,{\varvec{a}}} \right)\), which is obtained through CFD computations. It is iteratively computed during the processes by the optimizer and MCMC, due to variations of the parameters \({\varvec{a}}\) and \(\sigma\). The iteration number of this process could be the order of 103 to 106. It means that this order of times of the CFD computations are needed. This is the most time-consuming part in the whole process. The iteration number is unable to be predicted and this process itself is infeasible in practice especially for 3D aircraft test cases. A surrogate-assisted strategy is used for efficiency. A series of the following surrogate-assisted strategy is conducted in a framework of the DLR’s surrogate and reduced-order modeling toolbox called SMARTy. Recent articles on research works using SMARTy are found in Refs. [20, 21].

Once a surrogate model is constructed, no additional CFD computations are needed. A surrogate model is constructed at each flow condition (and/or at each spatial coordinate) \({\varvec{x}}^{\left( i \right)}\). The process of building a surrogate model is as follows. First, search domains of the turbulence model parameter \({\varvec{a}}\) are required to be decided. We set this as the ranges from 0.5 to 1.5 \({\varvec{a}}_{org}\) as already defined in Table 1, where \({\varvec{a}}_{org}\) is a vector of a set of the original turbulence model coefficients. Sampling locations where CFD computations are executed are determined by design of experiment (DoE). We used Sobol sequence, which is categorized as a kind of quasi Monte Carlo sampling methods [22, 23]. The Sobol sequence is able to keep the uniformity (the low-discrepancy) of the sampling filling the space even when the dimensionality of the input parameter is relatively high. It is also able to add sequentially low-discrepancy sampling points when the original sample size is not sufficient to achieve the accuracy of the constructed surrogate model. Based on the sample points generated by DoE, a surrogate model is constructed. Note that the process of constructing a surrogate model is theoretically the same as the statistical inference method in the previous subsection (a regression model by Eq. (12) or analytical solutions of the dual representation of Eq. (13) in e.g. Gaussian processes, where \({\varvec{x}}\) and \({\varvec{a}}\) in the previous subsection corresponds to \({\varvec{a}}\) and parameters to describe the surrogate model, respectively).

The constructed surrogate model eventually can be recognized as function approximation on the parameter space of \({\varvec{a}}\) at each flow condition (and/or at each spatial coordinate) as \(\tilde{y}^{\left( i \right)} = \tilde{f}\left( {{\varvec{a}};{\varvec{x}}^{\left( i \right)} } \right)\), where noted that \({\varvec{a}}\) is now input but \({\varvec{x}}^{\left( i \right)}\) as a set of parameters of the flow conditions is fixed. In practice, the new input of the surrogate model \(\tilde{f}\) corresponds to a new turbulence model parameter \({\varvec{a}}_{new}\) nominated during the process in the optimizer/MCMC. Thus, the output \(\tilde{f}\left( {{\varvec{a}}_{new} ;{\varvec{x}}^{\left( i \right)} } \right)\), which is iteratively computed in the parameter space \({\varvec{a}}\) during the optimization or MCMC process, is efficiently obtained via the surrogate model at each fixed flow condition (and/or at each spatial coordinate) \({\varvec{x}}^{\left( i \right)}\). Now the most time-consuming part has been replaced by the limited CFD computations on the DoE sample points, which are now predictive in terms of the cost. The required sample size to construct a surrogate model is normally dependent on the dimensionality of the input parameter \({\varvec{a}}\). The sample size is shown later at each test case in Sect. 3. The surrogate models here used are Gaussian processes with hyperparameter tuning for prediction of drag coefficient as \(f\), and additionally assisted by proper orthogonal decomposition (POD) for the cases of prediction of pressure coefficient. The kernel function commonly used in the Gaussian processes is a Gaussian kernel. The hyperparameter tuning as point-estimate to determine the probability distribution is achieved by the MLE. In prediction of pressure coefficients, multiple output of \({\mathbf{Y}} = \left( {y^{\left( 1 \right)} ,y^{\left( 2 \right)} , \ldots ,y^{\left( N \right)} } \right)\) is assisted by the POD at once for multiple input \({\mathbf{X}} = \left( {{\varvec{x}}^{\left( 1 \right)} ,{\varvec{x}}^{\left( 2 \right)} , \ldots ,{\varvec{x}}^{\left( N \right)} } \right)\), where \({\varvec{x}}\) is a spatial coordinate in this case, \(N\) as the sample size can be the number of the surface output of a wing.

During the entire process, a variance-based sensitivity analysis can be additionally inserted for the purpose of selection of dominant model coefficients to QoI as the output. The results of the sensitivity analysis are expected to bring physical aspects of the turbulence model coefficients. It can even bring dimensionality reduction, which can benefit of the computational costs. The required sample size for the DoE is basically dependent on the dimensionality of the input parameter space \({\varvec{a}}\). Therefore, it directly brings reduction of the required number of CFD computations. However, the sensitivity analysis itself requires the whole information using the original dimensionality of the parameter. Hence, it can be useful to use the results of a test case to other test cases to proceed the whole process with the selected coefficients. The global sensitivity of the input variables with respect to the output is computed by DoE samples using the Sobol sequence assisted by the constructed surrogate models. The execution of the Sobol indices itself was done by the Python library called SALib [24]. Details of the theory of the Sobol indices and sensitivity analysis methods in general can be found in Ref. [25].

Finally the flow chart of the data-driven method in this paper is shown in Fig. 2. This flow chart describes the direct process used in the application test cases in Sect. 3. The estimation of the modes by the optimizer can be a strategy for double check if the MCMC works properly.

Fig. 2
figure 2

Flow chart and simple sketch of surrogate-based approach

3 Applications and results

Two case studies were performed to validate and demonstrate the methods presented above: (a) a two-dimensional RAE2822 airfoil and (b) a three-dimensional full aircraft model based on the Airbus XRF1. Analyses of both test cases targeted the transonic regime. For the 2D test case, artificial experimental data was used to validate the proposed framework. Then, for the 3D test case, several sets of proprietary wind tunnel data for XRF1 were used to compare the simulation results against the experimental results. In this article, actual values cannot be published for XRF1 due to reasons of confidentiality. Therefore, quantitative results and/or relative evaluations are shown in Sect. 3.2. The 3D case is, however, valuable to demonstrate that the framework can be applied to real, industrial-grade test cases.

The parameter settings of the MCMC are common for all the test cases in this article. The sample size is 105 and the burn-in ratio is 0.8. The step length for \({\varvec{a}}\) normalized by \({\varvec{a}}_{org}\) and \(\sigma\) normalized by \(\hat{\sigma }\) are 0.025 and 0.05, respectively. Here \(\hat{\sigma }\) is the noise parameter obtained by the deterministic approach obtained by an optimizer (e.g. by Eq. (11)). Note just for visualization issues of the following figures that in the graphs of marginal posterior distributions of \({\varvec{a}}\) and \(\sigma\) obtained by the MCMC, the graphical ranges of \({\varvec{a}}\) and \(\sigma\) are common for all the figures in this article. The range of \({\varvec{a}}\) is from 0.5 to 1.5 \({\varvec{a}}_{org}\). The center is \({\varvec{a}}_{org}\). That of \(\sigma\) is from 0 to 5 \(\hat{\sigma }\) (only Fig. 7 is 0 to 2 \(\hat{\sigma }\)). This indicates that the mode of the posterior of \(\sigma\) should be located one fifth from the left (or the middle in Fig. 7) when the MCMC works well by producing the same results as the optimizer. Note also that all the predictions depicted by lines, and predictive distributions depicted by filled colored areas, have only their valid points/ranges at the discrete input flow conditions since the surrogate models are available for those discrete input flow conditions.

3.1 2D RAE2822 Airfoil

The RAE2822 airfoil is introduced here to validate the methodology and to investigate the characteristics of the turbulence model closure coefficients. The baseline Mach number is 0.734 and it varies in the transonic regime. The angle of attack and the Reynolds number are fixed at 2.79 degree and 6.5 million, respectively. The mesh used for CFD computations and a \(C_{p}\) visualization at this flow condition are shown in Fig. 3. The number of nodes is around 29 thousand.

Fig. 3
figure 3

RAE2822 airfoil configuration with grid overview and \(C_{p}\) visualization at Mach number of 0.734

For the purpose of dimensionality reduction, dominant closure coefficients, with respect to aerodynamic coefficients, are selected based on their Sobol indices. These selected closure coefficients are subsequently used as parameter \({\varvec{a}}\) for the XRF1 test case to reduce the computational costs; the assumption being that the flow characteristics are similar to each other in the transonic regime. Note that the results of the Sobol indices depends on the QoI as the output.

3.1.1 Sobol indices as preprocessing

The required information to generate the Sobol indices is obtained through surrogate models, which contain all of the closure coefficients as input parameters and the QoI as the output value. In this case the drag coefficient (\(C_{d}\)) was selected as the QoI to generate the Sobol indices. The surrogate models of \(C_{d}\) were constructed at each flow condition where the Mach number ranged from 0.704 to 0.784 in 0.01 increment. Each surrogate model was built using 100 DoE sample points generated between 0.5 and 1.5 \({\varvec{a}}_{org}\).

Table 2 shows results of the Sobol indices over the Mach number range. The percentages indicate the relative amount of contribution of the turbulence model coefficients to \(C_{d}\) at each flow condition. The mean values of the percentages at each coefficient are also shown in Table 2. It can be observed from the mean results that \(\kappa\) and \(c_{b1}\) are the most dominant followed by \(\sigma_{SA}\) and \(c_{v1} .\) This ranking of the result has similar characteristics to the works by Papadimitriou et al. [26] and Da Ronch et al. [2] when they investigated the original S–A. They listed \(\kappa\), \(c_{b1}\) and \(c_{v1}\) as the most influential coefficients to the output uncertainty of their quantities of interest. From the result in Table 2, the coefficients resulting from the negative S–A model which are not present in the original S–A model are not influential.

Table 2 Contribution percentage of the turbulence model coefficients to drag coefficient (\(C_{d}\)) at each flow condition

The selection process to determine which coefficients are dominant is similar to the mindset enforced when POD is employed and relies on using eigenvalues. Here we decided to truncate the list of dominant coefficients at 90%. Eventually the coefficients, \(\kappa\), \(c_{b1}\) and the \(\sigma_{SA}\), were selected as the dominant coefficients. These selected coefficients are set to the components of the turbulence model parameter \({\varvec{a}}\) as \({\varvec{a}} = \left( {\sigma_{SA} ,\kappa ,c_{b1} } \right)\) to be inferred in the following test cases. The process described so far is a preprocessing process and leads into the learning process using observed data. From now as a validation test of the framework, data-driven turbulence modeling results are shown by using artificial data.

3.1.2 Inference using drag coefficient

First the drag coefficient (\(C_{d}\)) is used as the QoI. Figure 4 shows marginal posterior distributions of each component of \({\varvec{a}}\) and \(\sigma\) after the learning process, and the obtained predictive distribution. As described in Sect. 2, the posterior distributions \(p\left( {{\varvec{a}},\sigma {|}{\mathbf{X}},{\mathbf{Y}}} \right)\) and the predictive distributions \(\left. {p\left( {y{|}{\varvec{x}}, {\mathbf{X}},{\mathbf{Y}}} \right)} \right|_{{\varvec{a}}}\) by Eq. (15) (contributed only by the closure model coefficient \({\varvec{a}}\)) were obtained directly by the MCMC sampling. Figure 4a also includes the result of \(\hat{\user2{a}}\) obtained through the optimization for comparison.

Fig. 4
figure 4

Results of predictions using \(C_{d}\). a Marginal distributions of a posterior of the turbulence model parameter (red dotted lines indicate \(\hat{\user2{a}}\)), b predicted \(C_{D}\) distributions by the point-estimate (green line), and the predictive distributions of \(C_{D}\) by the interval-estimate (light green fill as 95% CI), which are also shown in the right side of the figure as distributions. The dotted orange line is \(C_{D}\) distributions by the original model

The artificial data is also depicted in Fig. 4b. It was created based on CFD results obtained by using \({\varvec{a}}_{org}\) with independently and randomly generated noise from the Gaussian distribution \({\mathcal{N}}\left( {0,\sigma_{org}^{2} } \right)\), where \(\sigma_{org}\) is a constant value set to \(\sigma_{org}\) = 0.001 (10 drag counts), which was manually set by the authors. Since we know that the generated noise follows an independently identically distributed (i.i.d) Gaussian distribution, in this case, it is explicitly justified to use Eq. (10) as the likelihood function. It can be confirmed from Fig. 4a that \(\sigma_{SA}\) as \(a_{1}\) (the first element of \({\varvec{a}}\)) has greater uncertainty than the other coefficients in the relative ranges defined by \({\varvec{a}}_{org}\). In Fig. 4b, a reference result of the output obtained by \({\varvec{a}}_{org}\) and a result obtained by \(\hat{\user2{a}}\) (by the point-estimate) are depicted by the dotted orange lines and the green lines, respectively. The values of \(\hat{\user2{a}}\) and \(\hat{\sigma }\) obtained by the optimizer is \(\hat{\user2{a}}\) = [0.801, 0.417, 0.129] and \(\hat{\sigma }\) = 0.000689, respectively. Hence \(\hat{\user2{a}}\) obtained by the optimizer agrees well with the modes of the marginal posterior distributions obtained by the MCMC in Fig. 4a, However, the obtained \(\hat{\user2{a}}\) and \(\hat{\sigma }\) could not detect \({\varvec{a}}_{org}\) and \(\sigma_{org}\) as the reference (each component of \({\varvec{a}}_{org}\) is at the center of the horizontal axis). Nevertheless, it can be confirmed from Fig. 4a that the marginal posterior distributions are wide and covers \({\varvec{a}}_{org}\), \(\sigma_{org}\) with high probability density except for the first component of \({\varvec{a}}_{org}\). Therefore, a lack of the data for this amount of noise has been revealed by the results of the posterior obtained by the MCMC. It is emphasized that this information is never detected by the point-estimate.

Thus, this test case using the artificial data is not only validating the developed method but moreover offers further insight into underlying flow physics. For example, it can be observed that the influence of the turbulence closure coefficients is less when predicting the drag coefficient for a Mach number around 0.74 in contrast to higher and lower Mach numbers.

Next, the same test case but using prior information is also investigated. Prior can be set to any arbitrary probability of the parameters \({\varvec{a}}\) and \(\sigma\) independently, e.g. as \(p\left( {\varvec{a}} \right)\). So far, the constraints of the parameter as \({\varvec{a}} \in\)[0.5 \({\varvec{a}}_{org}\),1.5 \({\varvec{a}}_{org}\)] have been used in the process of the optimization and MCMC. Hence, these constraints are implicitly employed as prior information in the form of uniform distributions for all the test cases (see Appendix 1). In contrast, we enforce a prior distribution for \({\varvec{a}}\) which is Gaussian and can be represented as:

$${p\left( {\varvec{a}} \right) = {\mathcal{N}}\left( {{\varvec{a}}{|}{\varvec{a}}_{org} ,diag\left( {{\varvec{\sigma}}_{pri}^{2} } \right)} \right),}$$
(16)

where \({\varvec{\sigma}}_{pri} = \left( {\sigma_{pri,1} ,\sigma_{pri,2} , \cdots ,\sigma_{pri,M} } \right)\) and \(M\) is the dimensionality of the parameter \({\varvec{a}}\). Therefore it is the number of the turbulence model coefficients to be inferred. \(\sigma_{pri,j}\) indicates the variance of the \(j\) th element of the parameter \({\varvec{a}}\).

Figures 5 and 6 show marginal posterior distributions and predictive distributions, respectively, when the specific priors expressed by Eq. (16) are respected during the inference process. Priors for each element of the parameters can be independently set. Therefore, prior information is introduced at three different locations: (A) only for \({\varvec{a}}\), (B) only for \(\sigma\) and (C) for both \({\varvec{a}}\) and \(\sigma\). The function to be evaluated by the optimizer and the MCMC in case (A) is Eq. (16). Those for cases (B) and (C are \(p\left( {{\mathbf{Y}}{|}{\mathbf{X}},{\varvec{a}},\sigma } \right)p\left( {\varvec{a}} \right)\) and \(p\left( {{\mathbf{Y}}{|}{\mathbf{X}},{\varvec{a}},\sigma } \right)p\left( {\varvec{a}} \right)p\left( \sigma \right)\), respectively, where \(p\left( \sigma \right)\) can be obtained by as again following Eq. (16). The mode as (\(\hat{\user2{a}}\), \(\hat{\sigma }\)) obtained by direct optimization for each case is also described in Fig. 5.

Fig. 5
figure 5

Marginal distributions of a posterior of the turbulence model parameter \({\varvec{a}}\) when a prior is used (red dotted lines indicate \(\hat{\user2{a}}\))

Fig. 6
figure 6

Predicted \(C_{d}\) distributions when a prior is used, computed by a calibrated turbulence model parameter by the point-estimate, and the predictive distributions of \(C_{d}\) inferred by the interval-estimate

Due to the prior functions, the posteriors shown in Fig. 5 can be regarded as having updated from those in Fig. 4a. For example, in case (B), the uncertainty of \(\sigma\) as noise is reduced. In Fig. 5a, the uncertainty of \({\varvec{a}}\) is also reduced and the mode moved closer to \({\varvec{a}}_{org}\). Moreover, the resulting predictive distribution \(\left. {p\left( {y{|}{\varvec{x}}, {\mathbf{X}},{\mathbf{Y}}} \right)} \right|_{{\varvec{a}}}\) in Fig. 6a is smaller compared to Fig. 4b and is distributed around the reference result.

3.1.3 Inference using pressure coefficient

After using the global quantity \(C_{d}\) as the QoI, next the pressure coefficient (\(C_{p}\)) distributions is investigated. As for the drag coefficient analysis, artificial noise data is generated. Similar to before, the data was created based on CFD results using \({\varvec{a}}_{org}\) with independently and randomly generated noise from a Gaussian distribution \({\mathcal{N}}\left( {0,\sigma_{org}^{2} } \right)\), where \(\sigma_{org}\) = 0.1. The likelihood function is expressed as follows since the probabilistic model of the noise \(\varepsilon\) is assumed to be a Gaussian distribution \({\mathcal{N}}\left( {0,\sigma^{2} } \right)\) so that the probabilistic model of the output \(y\) that we defined is \({\mathcal{N}}\left( {f\left( {{\varvec{x}}^{\left( i \right)} ,{\varvec{a}}} \right),\sigma^{2} } \right)\). Therefore the likelihood function when the dataset \({\mathcal{D}} = \left( {{\mathbf{X}},{\mathbf{Y}}} \right)\) is:

$${p\left( {{\mathbf{Y}}{|}{\mathbf{X}},{\varvec{a}},\sigma } \right) = \mathop \prod \limits_{j = 1}^{{N_{data} }} \mathop \prod \limits_{i = 1}^{{N_{j} }} {\mathcal{N}}\left( {y^{{\left( {i,j} \right)}} {|}f\left( {{\varvec{x}}^{{\left( {i,j} \right)}} ,{\varvec{a}}} \right),\sigma^{2} } \right),}$$
(17)

where \(N_{data}\) and \(N_{j}\) indicate the number of data (equivalent to the number of the flow conditions) and the number of the available pressure coefficient at each flow condition, respectively. Note that this equation means that one unique variance \(\sigma^{2}\) to represent the output noise \(\varepsilon\) is assumed to be common at all the flow conditions, i.e. assuming that the noise level is the same at any flow conditions.

Figure 7 shows marginal posterior distributions and their predictive distributions \(\left. {p\left( {y{|}{\varvec{x}}, {\mathbf{X}},{\mathbf{Y}}} \right)} \right|_{{\varvec{a}}}\) at Mach numbers of 0.704, 0.734 and 0.764. The reference results obtained by \({\varvec{a}}_{org}\) and the results by the point-estimate are described as the dotted orange lines and the green lines, respectively. It can be seen that the shock waves around the mid-chord when the Mach number is at 0.704 and 0.734 can be accurately detected by the deterministic approach (the green line match the reference depicted by the orange dotted lines). The uncertainty shown as 95% credible interval (CI) is wider at the Mach number of 0.704. This uncertainty is the epistemic uncertainty associated with the turbulence closure coefficients. On the other hand, the point-estimate could not detect the reference at the Mach number of 0.764 but the resulting wide uncertainty (shown as 95% CI) by the interval-estimate covers the reference. These results mean that the point-estimate cannot detect the reference but the interval-estimate can as a probability and furthermore can indicate sensitive flow domains/conditions accordingly, which was not included in the given artificial data as information.

Fig. 7
figure 7

Results of predictions using \(C_{p}\) distributions of 3 flow conditions (Mach numbers of 0.704, 0.734 and 0.764) when no prior is used. a Marginal distributions of a posterior of the turbulence model parameter \({\varvec{a}}\) (red dotted lines indicate \(\hat{\user2{a}}\)), b predicted \(C_{p}\) distributions by the point-estimate (green line), and the predictive distributions of \(C_{p}\) by the interval-estimate (light blue fill as 95% CI). The dotted orange is the reference

3.2 XRF1 generic long-range transport aircraft

In this section the XRF1 is investigated relying on high-quality but proprietary wind tunnel data at transonic flow conditions and flight Reynolds numbers. XRF1 is an Airbus provided industrial standard multi-disciplinary research testcase representing a typical configuration for a long-range wide body aircraft. The XRF1 research testcase is used by Airbus to engage with external partners on development and demonstration of relevant capabilities / technologies.

The XRF1 wind tunnel data set at hand consists of drag (\(C_{D}\)) and pressure coefficients (\(C_{p}\)) at nine flow conditions, namely Case 1 to Case 9. The Mach number and the Reynolds number are fixed between the cases while the angle of attack (AoA) is varied. The Mach number and the Reynolds number are around 0.86 and 25 million, respectively. All the flight conditions are within the transonic regime and subsequently shock waves are present on the wing. In this test case, no specific prior is used (the improper prior is implicitly used) to learn the coefficients only from the wind tunnel data.

The inference is executed on two problems:

  1. 1.

    use \(C_{D}\) data from all the nine cases at once,

  2. 2.

    use \(C_{p}\) data available at three sections of the wing focusing on one Case only.

At each flow condition, the aircraft configuration is deformed due to its aeroelastic response. However, in this paper this influence is neglected and a fixed aircraft configuration is used for CFD computations for all nine cases. This fixed configuration has the surface deformation measured for Case 4. Hence for the second analysis focusing on \(C_{p}\) no deformation inconsistency is present. For the \(C_{D}\) analysis the inconsistency in surface deformations is expected to emerge to some extend as the variance \(\sigma^{2}\) in Eq. (9). Figure 8 shows the corresponding grid used for RANS computations. The number of grid nodes is around 6.3 million. As one reference showing a validation result of the grid (which also causes an epistemic uncertainty but evaluated in the noise parameter when it exists), we take an example that the difference of \(C_{D}\) between the corresponding data and a RANS computation using the original turbulence model coefficient \({\varvec{a}}_{org}\) is 3 drag counts (where one drag count is 10–4).

Fig. 8
figure 8

XRF1 configuration at Case 4 and grid overview

The results of the Sobol indices for the RAE2822 airfoil are used to downselect the turbulence model parameters which are inferred for the XRF1 assuming that both cases share underlying physics due to the transonic flow regime. Hence, the turbulence model coefficients \(\kappa\), \(c_{b1}\) and \(\sigma_{SA}\) are the inferred coefficients group in the parameter \({\varvec{a}}\) (\({\varvec{a}} = \left( {\sigma_{SA} ,\kappa ,c_{b1} } \right)\)). All neglected coefficients are fixed at the original S–A model values given in Table 1.

Surrogate models for the drag coefficient are constructed at each flow condition while for the pressure coefficient distributions one model for each wing section is computed. Since the dimensionality of the turbulence model parameter \({\varvec{a}}\) is 3, the required number of CFD computations has been reduced compared to the 12-dimensional airfoil case. A minimum of 70 DoE sample points to construct a surrogate model at each flow condition is used. This varying number of DoE samples is caused from diverging CFD simulations which occur once a combination of closure coefficients has been selected which violate the underlying RANS modeling assumptions.

3.2.1 Inference using drag coefficient

First, turbulence model coefficients are inferred based on the drag coefficient \(C_{D}\). The dataset used consisting of the input \({\mathbf{X}}\) and the output \({\mathbf{Y}}\) at all the nine flow conditions as \({\mathbf{X}} = \left( {{\varvec{x}}^{{\left( {case1} \right)}} ,{\varvec{x}}^{{\left( {case2} \right)}} , \cdots ,{\varvec{x}}^{{\left( {case9} \right)}} } \right)\) and \({\mathbf{Y}} = \left( {{\varvec{y}}^{{\left( {case1} \right)}} ,{\varvec{y}}^{{\left( {case2} \right)}} , \cdots ,{\varvec{y}}^{{\left( {case9} \right)}} } \right)\), where \({\varvec{x}}\) and \({\varvec{y}}\) are the AoA and \(C_{D}\), respectively. The experimental data for \(C_{D}\) is shown in Fig. 9. As aforementioned, the deformation used for all computations is fixed and belongs to Case 4 which is the case of the second lowest angle of attack. Figure 9 also depicts the possible \(C_{D}\) region in the search space \({\varvec{a}} \in\)[0.5 \({\varvec{a}}_{org}\),1.5 \({\varvec{a}}_{org}\)]. This region is spanned by directly using all available DoE sample pointsand plotting their minimum and maximum values. It has to be noted that this coverage does not guarantee that a possible combination of parameters exists for each \(C_{D}\) value inside. Nevertheless, it can be seen that all experimental data points are within the covered region. The \(C_{D}\) characteristics using the original coefficients \({\varvec{a}}_{org}\) are also described in Fig. 9.

Fig. 9
figure 9

Drag coefficient distribution of experimental data, CFD solutions using the original coefficients, and feasible CFD solutions range spanned by DoE samples

Results obtained by the point-estimate and the interval-estimate are shown in Fig. 10. Since the probabilistic model is exactly the same as that in the case of RAE2822 in the previous sub-section, the likelihood function used for the learning process is Eq. (10) as well. First of all, the results of the point-estimate are discussed. After the learning process, the parameters \({\varvec{a}}\) and \(\sigma\) in Eq. (10) are calibrated as \(\hat{\user2{a}}\) and \(\hat{\sigma }\). Figure 10 shows \(C_{D}\) characteristics over the nine angles of attack. The green line indicates the results computed by using the calibrated turbulence model parameter \(\hat{\user2{a}}\). At the angle of attack of Case 4, the predicted \(C_{D}\) by the calibrated parameter \(\hat{\user2{a}}\) (\(C_{D}^{calib}\)) corresponds well to the underlying data. For the other cases, \(C_{D}^{calib}\) is between the data \({\mathbf{Y}}\) and \(C_{D}^{org}\) at each angle of attack, which is the qualitative information obtained by the point-estimate. The difference between \(C_{D}^{calib}\) and \({\mathbf{Y}}\) is treated as the noise produced by a probability with the calibrated parameter \(\hat{\sigma }\). This noise represents all of the uncertainties except the uncertainty associated with the model parameter \({\varvec{a}}\). By following the model definition in sect. 2.2, the inaccuracy of the configuration due to aeroelasticity is also treated as the noise.

Fig. 10
figure 10

Results of predictions using \(C_{D}\) by the experimental dataset at nine flow conditions. a Marginal distributions of a posterior of the turbulence model parameter (red dotted lines indicate \(\hat{\user2{a}}\)), b predicted \(C_{D}\) distributions by the point-estimate (green line), and the predictive distributions of \(C_{D}\) by the interval-estimate (light green fill as 95% CI), which are also shown in the right side of the figure as distributions. The dotted orange line is \(C_{D}\) distributions by the original model

When following a stochastic approach, the marginal posterior distributions shown in the upper part of Fig. 10 have large uncertainties especially on \(a_{1}\) (= \(\sigma_{SA}\)) since it is fully distributed in the constrained range of [0.5 \({\varvec{a}}_{org}\), 1.5 \({\varvec{a}}_{org}\)]. The mode is around the lower bound. The noise parameter \(\sigma\) is also widely distributed and spreads around four times of the mode \(\hat{\sigma }\). The predictive distributions \(\left. {p\left( {y{|}{\varvec{x}}, {\mathbf{X}},{\mathbf{Y}}} \right)} \right|_{{\varvec{a}}}\) are depicted in the lower part of Fig. 10. It can be observed that the 95% CI at the highest AoA is quite wide compared to lower AoA. The predictive distribution at the second highest AoA is relatively narrow and the data point at this condition is not within the 95% CI. The noise parameter \(\sigma\) is actually estimated quite high even though the exact drag counts of \(\sigma\) cannot be described here. Therefore, the predictive distribution considering all the uncertainties computed by Eq. (13), should contain the data points. However, in this particular application case, there exists external noise due to the geometry difference which falsifies this assumption.

3.2.2 Inference using pressure coefficient

As described in the 2D airfoil case, at each flow condition, the amount of the information when using \(C_{p}\) distributions is much greater than when using \(C_{D}\). However, for practically relevant cases, \(C_{p}\) measurements are usually available only at several sections/locations. Figure 11 shows \(C_{p}\) distributions of the experimental data and feasible CFD solutions for Case 4. Three sections are available and each of them cover the whole surface of the section where \(C_{p}\) data is measured. namely Section \(i\) to Section \(iii\) ordered from wing root to tip. Since the angle of attack at Case 4 is negative, shock waves are generated on the lower surface of the wing.

Fig. 11
figure 11

Pressure distributions of experimental data and feasible CFD solutions by DoE samples at three sections

As the assumption in the inference process introduced in Sect. 2.2, the turbulence model parameter \({\varvec{a}}\) is stochastic. It is neither a spatio-temporal parameter like a function of the time and space nor a function of flow conditions. In definition of probabilistic models, we introduce two derivations (A) and (B) from two hypotheses:

  1. A.

    the same noise parameter at each section as one \(\sigma\)

  2. B.

    different noise parameters at each section as \(\sigma^{\left( j \right)}\), where \(j\) = 1,2,3

The likelihood function in the hypothesis (A) is the same as Eq. (17) in the 2D RAE2822 test case as follows:

$${p\left( {{\mathbf{Y}}{|}{\mathbf{X}},{\varvec{a}},\sigma } \right) = \mathop \prod \limits_{j = 1}^{{N_{sec} }} \mathop \prod \limits_{i = 1}^{{N_{j} }} {\mathcal{N}}\left( {y^{{\left( {i,j} \right)}} {|}f\left( {{\varvec{x}}^{{\left( {i,j} \right)}} ,{\varvec{a}}} \right),\sigma^{2} } \right),}$$
(18)

where \(x^{{\left( {i,j} \right)}}\) indicates in this case the spatial coordinate of \(i\) th point at the section \(j\), and \(N_{sec}\) indicates the number of the sections. Here we use the three sections, therefore \(N_{sec}\) = 3. A common noise parameter denoted by \(\sigma\) is set for all the sections.

On the other hand, the noise parameter of the likelihood function in the hypothesis B) can be expressed by a vector. Each element of the parameter vector consists of a different noise parameter \(\sigma^{\left( j \right)}\) at each section \(j\). The likelihood function is slightly modified from Eq. (18) as:

$${p\left( {{\mathbf{Y}}{|}{\mathbf{X}},{\varvec{a}},{\varvec{\sigma}}} \right) = \mathop \prod \limits_{j = 1}^{{N_{sec} }} \mathop \prod \limits_{i = 1}^{{N_{j} }} {\mathcal{N}}\left( {y^{{\left( {i,j} \right)}} \left| {f\left( {{\varvec{x}}^{{\left( {i,j} \right)}} ,{\varvec{a}}} \right),\sigma^{{\left( j \right)^{2} }} } \right.} \right),}$$
(19)

where \({\varvec{\sigma}} = \left( {\sigma^{\left( 1 \right)} ,\sigma^{\left( 2 \right)} , \ldots ,\sigma^{{\left( {N_{sec} } \right)}} } \right)\). The probabilistic models of the hypotheses (A) and (B) are summarized in Appendix 4 for further descriptions of the hypotheses behind.

The point-estimate and the interval-estimate are conducted on Eqs. (18) and (19) in order to obtain \(\hat{\user2{a}}\), \(\hat{\user2{\sigma }}\) (by the point-estimate), and \(p\left( {{\varvec{a}},{\varvec{\sigma}}} \right)\) (by the interval-estimate), where \(\hat{\user2{\sigma }} = \left( {\hat{\sigma }} \right)\) in Eq. (18), \(\hat{\user2{\sigma }} = \left( {\hat{\sigma }^{\left( 1 \right)} ,\hat{\sigma }^{\left( 2 \right)} , \cdots ,\hat{\sigma }^{{\left( {N_{sec} } \right)}} } \right)\) in Eq. (19), respectively.

Figure 12 shows the marginal posterior distributions and predictive distributions \(\left. {p\left( {y{|}{\varvec{x}}, {\mathbf{X}},{\mathbf{Y}}} \right)} \right|_{{\varvec{a}}}\) when a unique noise parameter is introduced. In the predictive distributions at all three sections, the epistemic uncertainties associated with the parameter \({\varvec{a}}\) (shown as 95% CI in the figure) are slightly wider around the shock locations and downstream thereof compared to more benign flow phenomena. However, they are so small that they do not cover the experimental data. In fact, the predictive distribution at section \(i\) even diverged further from the experimental data than the CFD prediction relying on the original model parameters \({\varvec{a}}_{org}\). These results are mainly caused by introducing the same noise parameter for all the sections.

Fig. 12
figure 12

Results of predictions using \(C_{p}\) distributions by the experimental dataset at one flow condition (Case 4) when the turbulence model parameter \({\varvec{a}}\) and the noise parameter \(\sigma\) are set to be common at all the three sections. a Marginal distributions of a posterior of the turbulence model parameter (red dotted lines indicate \(\hat{\user2{a}}\)), b predictions by the point-estimate and predictive distributions by the interval-estimate (showing 95% CI)

On the hypothesis B using Eq. (19), a more flexible assumption is used in which noise at each section is expressed independently. Figure 13 shows the marginal posterior distributions and predictive distributions \(\left. {p\left( {y{|}{\varvec{x}}, {\mathbf{X}},{\mathbf{Y}}} \right)} \right|_{{\varvec{a}}}\). The overall characteristics are similar to those of the hypothesis (A). However, the predictive distribution covers the shock wave at Section \(i\) slightly better than hypothesis A). Hence, assuming different noise parameters for each section seems more promising than assuming a common noise parameter. However, the results in Fig. 13 also clearly indicate that there is no possibility for the turbulence model parameter to properly predict all flow phenomena at the three sections by adjusting just three turbulence model parameters and aiming for a global model description.

Fig. 13
figure 13

Results of predictions using \(C_{p}\) distributions by the experimental dataset at one flow condition (Case 4) when the turbulence model parameter \({\varvec{a}}\) is set to be common at all the three sections but the noise parameter \(\sigma\) is set individually. Upper: marginal distributions of a posterior of the turbulence model parameter (red dotted lines indicate \(\hat{\user2{a}}\)). Lower: predictions by the point-estimate and predictive distributions by the interval-estimate (showing 95% CI)

Based on the previous results, the flexibility is further increased by allowing different turbulence model parameters \({\varvec{a}}\) for each section as \({\varvec{a}}^{\left( j \right)}\), where \(j\) = 1,2,⋯,\({ }N_{sec}\) (named hypothesis (C)). The likelihood function at each section is identical to Eq. (10). Figure 14 shows three sets of marginal posterior distributions on the three sections and the respective predictive distributions \(\left. {p\left( {y{|}{\varvec{x}}, {\mathbf{X}},{\mathbf{Y}}} \right)} \right|_{{\varvec{a}}}\). It can be observed in Fig. 14 that the shock wave at each section can be better detected by independently inferred parameter \({\varvec{a}}^{\left( j \right)}\). As before, the uncertainties associated with the individual parameters \({\varvec{a}}\) are larger around the shock locations. This showcases that the developed framework is able to detect location of increased physical complexity, e.g. shocks, and highlights them by a larger local uncertainty.

Fig. 14
figure 14

Results of predictions using \(C_{p}\) distributions by the experimental dataset at one flow condition (Case 4) when both the turbulence model parameter \({\varvec{a}}\) and the noise parameter \(\sigma\) are set independently at each section. a Marginal distributions of a posterior of the turbulence model parameter (red dotted lines indicate \(\hat{\user2{a}}\)), b predictions by the point-estimate and predictive distributions by the interval-estimate (showing 95% CI)

A natural extension of the previous strategy is the assumption that the parameter is spatially varying in general. For example, Duraisamy et al. [7] proposed a model correction term and then inferred it employing a learning process. Since the spatially varying parameter has the same dimensionality than the underlying computational grid, computations of a posterior by MCMC are infeasible and they are achieved by the other approximation methods based on definition of posterior distributions such as the variational inference and the Laplace approximation (see Appendix 2).

4 Conclusions

A framework for data-driven turbulence modeling under uncertainties was developed and applied to a 2D airfoil and an industrially-relevant 3D full aircraft case. The objective of this framework is to quantify the uncertainty of the closure coefficients and that of the quantities of interest associated with the closure coefficients by employing a Bayesian approach. In this paper, the closure coefficients of the Spalart Allmaras turbulence model were inferred. The partial availability of information, which is common for cases of industrial relevance, is regarded as an epistemic uncertainty that is addressed by a fully Bayesian approach, with special treatment for computing the predictive distribution. The resulting uncertainties can provide an insight into one of the model uncertainties due to the closure coefficients for predicting the quantities of interest by distinguishing it from other external uncertainties. In this work, these other external uncertainties are treated as noise. The other possible epistemic uncertainties in the CFD simulations such as the grid convergence were eliminated as much as possible in advance though they are classified into the noise if they exist. The uncertainties associated by all the parameters, which are the closure coefficients and the noise parameter, in the fully Bayesian model were rigorously evaluated by a Markov-Chain Monte Carlo (MCMC) sampling technique. To overcome the bottleneck of high computational cost in the MCMC process associated with repetitive RANS evaluation, a surrogate-based methodology has been employed.

The 2D airfoil test case with artificial data has been used to validate the methodologies integrated in the framework. Moreover, dominant closure coefficients were selected by using Sobol indices as a preprocessing step that further increased the efficiency of the surrogate-based approximation. It was shown that known model coefficients can be rediscovered by relying on data. Furthermore, it was confirmed not only that the closure coefficients and the artificial noise were replicated by the developed method but also that the epistemic uncertainty associated with the closure coefficients are large around the shock locations. What is important in this replication test case is that the information of this epistemic uncertainty was not specified in the problem setting but was derived by the developed method.

For the 3D aircraft case the information obtained from the airfoil Sobol indices analysis has been reused to keep computational cost at a feasible level. Inference of coefficient by analyzing pressure coefficients showed that the epistemic uncertainty associated by the coefficients are highly detected in regions of sensitive flow, such as the shock position and downstream thereof. It was also shown that the uncertainty associated by the noise is quite large compared with that by the coefficients in this test case.

Obtained results, and in particular the epistemic uncertainties associated by the coefficients, can be used for various applications. For example, one option could be to feedback information to the experimental measurement strategies to determine at which locations on the wing new measurement devices should be placed. Another opportunity is the propagation of uncertainties towards unobservable flow conditions with the assistance of surrogate models and the posterior of the closure coefficients to enhance the modeling of e.g. loads at the edge of the flight envelope. The developed framework is flexible in terms of modelling, i.e. other closure models can be applied as well as other probabilistic models.