49th AIAA Aerospace Sciences Meeting including the New Horizons Forum and Aerospace Exposition 4 - 7 January 2011, Orlando, Florida

AIAA 2011-660

Uncertainty quantification for laminar-turbulent transition prediction in RANS turbomachinery applications Ren´e Peˇcnik∗, Jeroen A.S. Witteveen†, and Gianluca Iaccarino‡ Stanford University, Stanford CA, 94305 The effect of physical variability and model uncertainty on laminar-turbulent transition in transonic gas turbine compressors is computed using Stochastic Collocation methods. Aleatoric and epistemic uncertainties are considered in an adiabatic flat plate validation and turbine guide vane simulations. The computational results show that the variability has significant impact on the transition location for the turbine guide vane simulations and, consequently, on the reliability of the model. The model uncertainty accounts to a large extent for the difference between the deterministic simulation and the experiments. The results from the Simplex Stochastic Collocation method show a more robust convergence than those of Stochastic Collocation based on Clenshaw–Curtis quadrature.

I.

Introduction

turbomachines and especially in aircraft engines the Reynolds numbers that determine the evolution of boundary layers are relatively low, hence a large part of the flow along the blade surfaces is laminar Iornthetransitional. Bypass transition is the dominant form of transition in turbomachinery due to the high turbulence levels, e.g., generated by upstream blade rows. The boundary layer development, losses, efficiency, and heat transfer are greatly affected by the laminar-to-turbulent transition. Therefore, the ability to accurately predict the transition process is crucial for the design of efficient and reliable machines.8 Considerable effort has been spent in the past two decades to develop transition models for engineering applications to predict transitional boundary layers for various kinds of flows. In general, these models rely entirely on empirical correlations obtained from existing data sets for simple flow configurations. Unfortunately, for complex flows there is only limited experimental data available that can be incorporated to calibrate these models. Furthermore, in complex flow configurations the lack of knowledge of the exact inlet boundary conditions adds a second type of uncertainty. Previous uncertainty analysis studies have shown that transonic gas turbine compressors are sensitive to these uncertainties.4, 6 Therefore, this work focuses on the quantification of uncertainties due to the lack of knowledge of the physical processes associated with boundary layer transition (also called epistemic uncertainty or model form uncertainty) and the uncertainties of turbulence inlet conditions for complex flow situations (also called aleatory or irreducible uncertainties associated with boundary conditions in general). The epistemic uncertainties as well as the aleatory uncertainties are identified and addressed in the simulations of a transonic turbine guide vane. The flow is modeled using the Reynolds-Averaged Navier-Stokes (RANS) equations closed with the  θt k − ω shear stress transport (SST) turbulence model7 and the transition is modeled using the γ − Re 5 correlation based transition model of Langtry and Menter. The correct implementation of these models is first validated for commonly used transition benchmark cases of a flow over an adiabatic flat plate, with and without imposed pressure gradient.13 The models are then applied to the boundary layer transition on the ∗ Postdoctoral Fellow, Center for Turbulence Research, Building 500, Stanford University, Stanford, CA 94305-3035, email: [email protected] † Postdoctoral Fellow, Center for Turbulence Research, Building 500, Stanford University, Stanford, CA 94305-3035, email: [email protected] ‡ Assistant Professor, Mechanical Engineering Dept., Building 500, Stanford University, Stanford, CA 94305-3035

1 of 14 American Institute Aeronautics Copyright © 2011 by the American Institute of Aeronautics and Astronautics, Inc. All of rights reserved. and Astronautics

transonic turbine guide vane of the VKI test cases MUR241 and MUR235.1 The Uncertainty Quantification (UQ) method used is Stochastic Collocation2, 18 (SC) with Clenshaw-Curtis (CC) quadrature points. It propagates the input uncertainties in the inlet turbulence intensity T uin , turbulence Reynolds number ReT , and the uncertainties associated with the transition model through the computational domain. The results for the case with three uncertainties in ReT and two epistemic parameters in the transition modeling are compared to those of the more robust Simplex Stochastic Collocation16, 17 (SSC) method. The paper is organized as follows. Sections II and III give a brief overview of the flow solver and the models used in this work. Furthermore, in Section III an approach to quantify the uncertainties associated with the physical processes in boundary layer transition is proposed, which is particularly important for transonic flows. The SC and SSC methods employed for the uncertainty propagation are discussed in Section IV. After the validation for the flat plate in Section V, the uncertainty analysis of the VKI turbine vane is considered in Section VI and VII. The main conclusions are summarized in Section VIII.

II.

Reynolds-averaged Navier-Stokes solver

The computations were performed using the in-house Reynolds-averaged Navier-Stokes (RANS) solver, developed at the Center for Turbulence Research at Stanford University. The flow solver is a parallel solver for the solution of the compressible Navier-Stokes equations on unstructured meshes based on a finite volume formulation and implicit time-integration on arbitrary polyhedral mesh elements.9, 10 The governing equations are written in conservative form as   ∂ UdΩ + [F (U) − Fv (U)] dA = 0 , (1) ∂t Ω ∂Ω where U = U(x, t) is the state variable, F(U) and Fv (U) are the convective and viscous fluxes, respectively, and Ω and ∂Ω are the physical domain of interest and its boundary. In particular, we consider U F(U) F(U)v

=

[ρ, ρv, E]T

= =

[n · ρv, v (n · ρv) +pn, (E + p) (v · n)] [0, n · Π, v · (n · Π) + n · Q]T ,

(2) T

where ρ, v, p, E, Π, Q, n represent density, Cartesian velocity vector, pressure, total energy, stress tensor, heat flux vector and outward pointing unit vector normal to the surface, respectively. The code is entirely written in C++ and uses subdomain decomposition and the message passing interface (MPI) as the parallel infrastructure. The flow quantities are integrated in conservative form using a Newton-Rhapson implicit scheme:   I ∂R 1  + [F (U) − Fv (U)] Af . (3) ΔU = −R (Un ) , with: R (U) = Δt ∂U V f

A Taylor expansion is used to formulate the Jacobian matrices ∂R/∂U for the inviscid and viscous fluxes. The inviscid fluxes are computed with the HLLC approximate Riemann solver3, 15 based on state values reconstructed at the face center using a spatial second order approximation. The viscous fluxes are approximated with a central difference approximation, likewise leading to a second order accurate scheme. The resulting large sparse system (the Jacobian matrices for the inviscid fluxes are obtained using first-order discretization) is solved with the generalized minimal residual method (GMRES) using the freely available linear solver package PETSc.12 The scalar transport equations for the turbulence and transition model are solved segregated after each pseudo time step for the Navier-Stokes equations. The transport equation for a generic scalar φ can be written in conservative form as    ∂ (ρφ) dΩ + [φ (n · ρv) − Fv (φ)] dA = S (φ) dΩ , (4) ∂t Ω ∂Ω Ω where S (φ) is the scalar source term and Fv (φ) is the viscous flux for the scalar considered.

2 of 14 American Institute of Aeronautics and Astronautics

III.

Laminar-turbulent transition modeling

 θt correlation In order to account for laminar to turbulent transition in the test cases considered, the γ− Re based transition model of Langtry and Menter5 is used. It is a framework for implementing empirical transition criterions in general-purpose flow solvers that can be used in an unstructured and parallel code. The model solves two transport equations, namely the intermittency factor γ that describes the fraction of time in which the flow at at a certain location is turbulent, and the momentum thickness Reynolds number  θt that is used to trigger the transition onset. The intermittency is one in the free stream and in the fully Re developed turbulent boundary layer and zero in the pre-transitional laminar boundary layer. III.A.

Transition model

The transport equations in differential form are given by

   ∂ρui γ ∂ργ ∂ μt ∂γ + = Pγ − Eγ + μ+ , ∂t ∂xi ∂xi σf ∂xi

 θt  θt  θt ∂ρRe ∂ρui Re ∂ ∂ Re + = Pθt + σθt (μ + μt ) , ∂t ∂xi ∂xi ∂xi

(5)

(6)

where Pγ is the production and Eγ is the destruction of the intermittency, respectively. The term Pθt in (6) is the production of the momentum thickness Reynolds number and is given by ρ  θt (1 − Fθt ) , Reθt − Re (7) Pθt = cθt t with the empirical correlation criterion for the transition onset Reθt . The model is applicable to incompressible transitional flows. In transonic flows, compressibility and the presence of shock waves affect the onset and extend of transition considerably, and should be incorporated into the model. Unfortunately, there is only limited open literature about these compressibility effects on boundary layer transition. Qualitatively, increasing the Mach number results in a longer transition zone and a delay of the onset. However, a large discrepancy among different correlations, proposed by various groups, exists when this effect is quantified. These studies investigated the Mach number influence on the growth rate of the turbulent spots in the transitional boundary layer and the various results showed a decrease of the spot growth rate at Mach=1 by a factor of 1.16 to almost 6.14 Similar discrepancies exist regarding the Mach number influence on the transition onset. Instead of developing a correlation for the compressibility effect to match a particular experimental data set, this work focuses on quantifying the above mentioned uncertainties. Initially, given correlations are used and then utilized to assess the lack of knowledge, respectively the large uncertainties associated with the compressibility. Effectively, the decrease of the turbulent spot growth with increasing Mach number can  θt model by changing the intermittency production term Pγ . Steelant and be incorporated into the γ − Re Dick14 found that the intermittency production Pγ is proportional to the square root of the turbulent spot growth parameter, hence Pγ can by modified by

(8) Pγ∗ = (f (M ais ))Pγ with f (M ais ) as a function of the isentropic Mach number M ais . A simple linear function is constructed and given by: f (M ais ) = 1 − (0.14 + 0.7UP ) M ais , (9) with UP , a uniformly distributed random variable on the domain [0; 1] to assess the uncertainty of the spot growth parameter mentioned above. For M ais = 1 a decrease of 16% to 86% can be achieved and for M ais = 0 the function returns 1 and therefore does not change Pγ for incompressible flows. Similarly, the transition onset delay for transonic flows is incorporated by changing Reθt in (7) as a function of M ais : Re∗θt = (1 + 0.6UReθt M ais ) Reθt ,

(10)

with UReθt a uniformly distributed random variable on the domain [0; 1]. For M ais = 0 the transition onset criterion Reθt returns its incompressible value, whereas for M ais = 1 a maximum increase of 60% can be achieved. 3 of 14 American Institute of Aeronautics and Astronautics

III.B.

Underlying turbulence model

 θt transport equations use the k − ω shear stress transport (SST) turbulence model7 as underlying The γ − Re model to provide the turbulent quantities k − ω and the eddy viscosity μt for the Navier-Stokes solver. As the SST model has gained an increasing popularity during the last years a variety of modifications to various limiters within the model have been introduced; hence, different versions are available. The most important limiter functions of the model version used throughout this work are given below. The limiter for the eddy viscosity is given by: ρa1 k , (11) μt = max (a1 ω, ΩF2 ) which uses the magnitude of the vorticity Ω in the denominator of the limiter. The production limiter, that ensures realizable turbulent Reynolds stress tensors is: Pk∗ = min (Pk , 20Cμ kω) .

IV. IV.A.

(12)

Uncertainty quantification methods

Stochastic Collocation

The Stochastic Collocation (SC) method2, 18 is used to compute the effect of the input and model uncertainties on the laminar-turbulent transition in the turbomachinery application. The main idea behind Stochastic Collocation is to sample the quantity of interest at particular points in the parameter space. Integral statistics such as mean and standard deviation are then computed using quadrature or cubature rules. Non–integral statistics in the form of probability density functions are obtained from the Monte Carlo samples of the Lagrange interpolation of the quantity of interest over the sample points. In the case of a one–dimensional parameter space, these points are quadrature abscissas selected according to the probability measure of the random variable. Tensor product and sparse grid constructions are then utilized to generate the multi– dimensional abscissas. The choice of one–dimensional abscissas defines the properties of the interpolation formula or the integration rule. Common choices of these nodes are the Clenshaw–Curtis (CC) and Gaussian abscissas. The Clenshaw–Curtis abscissas are the extrema of the Chebyshev polynomials in the interval [−1, 1], see Fig. 1 for a two–dimensional example. For any choice of np > 1, these points are given by

Figure 1. Two–dimensional Stochastic Collocation Clenshaw–Curtis points.

 y = − cos j

π(j − 1) np − 1

 ,

j = 1, . . . , np ,

(13)

which renders to a nested rule in the sense that the set of lower order quadratures abscissas for np = 2i + 1 is a subset of that of a higher order one with np = 2j + 1 for integer values i < j. We adopt the CC rule in this work, since this hierarchical sampling property allows for estimating the quadrature error by comparing the results of different collocation orders. The response statistics are then computed as follows. Let f (y1 , . . . , yd )

4 of 14 American Institute of Aeronautics and Astronautics

be the quantity of interest as a function of independent random variables {y1 , . . . , yd } each with a probability density function ρk : Ω → Γk ⊂ R, k = 1, . . . , d, and let npd

np

Λ := {y11 , . . . , y1 1 } × · · · × {yd1 , . . . , yd

},

(14)

denote the set of multi–dimensional abscissas constructed by taking the tensor product of one–dimensional abscissas corresponding to each random variable yk . The integral–valued statistics of f can be approximated using the multi–dimensional quadrature integration formula   d  ··· g(f )(y1 , . . . , yd ) ρk (yk )dy1 · · · dyd E[g(f )] = ≈

Γ1

Γd

np1

npd





···

j1 =1

k=1

jd =1

g(f )(y1j1 , . . . , ydjd )w1j1 · · · wdjd ,

(15)

where wkjk is the weight associated with the quadrature point ykjk and g is a function defining the desired statistics of f . Clearly, the accuracy of the approximation (15) depends on the location ykjk of each abscissas as well as the total number of points npk along each direction k. Notice that cardinality of Λ increases exponentially fast with respect to the number of random variables d and npk . The tensor product collocation scheme is therefore not computationally efficient for systems with large number of random inputs. Alternatively, one can first construct an interpolant of f from samples in Λ and then estimate the integral and non–integral statistics of f using Monte Carlo samples of the interpolant. For this purpose, typically, a multi–dimensional Lagrange interpolation is adopted. More specifically, let npk

lkjk (yk ) =



ik =1, ik =jk

yk − ykik

ykjk − ykik

,

(16)

represent the one–dimensional Lagrange polynomials corresponding to abscissa ykjk , then the interpolant I(f ) of f is given by I(f )(y1 , . . . , yd ) =

np1  j1 =1

···

npd  jd =1

f (y1j1 , . . . , ydjd )(l1j1 ⊗ · · · ⊗ ldjd )(y1 , . . . , yd ),

(17)

where ⊗ denotes the Kronecker product. The desired statistics are approximated from samples of I(f ), for example, the expected value is computed as E[g(f )] ≈

N 1  g(f )(y1i , . . . , ydi ), N i=1

(18)

where y1i , . . . , yki , i = 1, . . . , N , are independent and identically distributed samples of y1 , . . . , yd . IV.B.

Simplex Stochastic Collocation

The global polynomial approximation of Stochastic Collocation can, however, be unreliable in case of large gradients in the response surface. This situation can, for example, occur for the heat flux in the transition region when the transition location is impacted by uncertainty. Also the spectral convergence of the Stochastic Collocation method can significantly reduce with an increasing number of uncertainties, due to the structured grid of the quadrature points in multiple random dimensions. Here, the Simplex Stochastic Collocation (SSC) method16, 17 is presented that combines the accuracy of polynomial interpolation with the robustness and the effectiveness in higher dimensions of Monte Carlo simulation based on random sampling. SSC discretizes the parameter space Ξusing non-overlapping simplex ne elements Ξj from a Delaunay triangulation of sampling points, with Ξ = j=1 Ξj and ne the number of elements. In each of the simplexes Ξj , the response surface of the quantity of interest u(ξ) as function of the random parameters ξ ∈ Ξ is approximated by a polynomial wj (ξ) wj (ξ) =

P 

cj,m Ψj,m (ξ),

m=0

5 of 14 American Institute of Aeronautics and Astronautics

(19)

with P + 1 coefficients cj,m and basis polynomials Ψj,m (ξ). The polynomials are found by the interpolation of the samples vk = u(ξ k ) at the vertexes ξ k of the simplex elements, with k = 1, . . . , ns and ns the number of samples. For higher degree interpolation, a stencil of sampling points vkj,l in the vertexes ξkj,l of surrounding simplexes is constructed, with l = 0, . . . , N and kj,l ∈ {1, . . . , ns }. The polynomial coefficients cj,m are then given by ⎡ ⎞ ⎛ ⎞ ⎤⎛ Ψj,0 (ξ kj,0 ) Ψj,1 (ξ kj,0 ) · · · Ψj,P (ξ kj,0 ) vkj,0 cj,0 ⎢ ⎟ ⎜ ⎟ ⎥⎜ ⎢ Ψj,0 (ξ kj,1 ) Ψj,1 (ξ kj,1 ) · · · Ψj,P (ξ kj,1 ) ⎥ ⎜ cj,1 ⎟ ⎜ vkj,1 ⎟ ⎢ ⎥⎜ . ⎟ = ⎜ . ⎟, (20) .. .. .. ⎢ ⎥⎜ . ⎟ ⎜ . ⎟ .. . ⎣ ⎦⎝ . ⎠ ⎝ . ⎠ . . . vkj,N cj,P Ψj,0 (ξ kj,N ) Ψj,1 (ξ kj,N ) · · · Ψj,P (ξ kj,N ) with N ≥ P . The robustness of the approximation is guaranteed by using a limiter approach for the local polynomial degree pj , based on the extension of the Local Extremum Diminishing (LED) concept to probability space. This ensures that no overshoots are present in the response interpolation in each of the elements Ξj min(wj (ξ)) ≥ min(u(ξ)) ∧ max(wj (ξ)) ≤ max(u(ξ)), (21) Ξj

Ξj

ΞJ

Ξj

for j = 1, . . . , ne . The initial samples consist of the extrema of the parameter ranges and one at the nominal conditions, see Fig. 2(a) for a two-dimensional example. The discretization is adaptively refined by calculating a refinement measure based on a local error estimate in each of the simplex elements. A new sampling point is then added randomly in the simplex with the highest measure and the Delaunay triangulation is updated. The sample is confined to a subdomain of the simplex to ensure a good spread of the sampling points, see Fig. 2(a). The refinement to ns = 17 samples, shown in Figure 2(b), leads to a superlinear convergence by increasing the polynomial degree pj with the increasing number of available samples ns . The sampling procedure is stopped when a global error estimate reaches an accuracy threshold. 1 2

0.8

random parameter ξ

random parameter ξ

2

1

0.6 0.4 0.2 0 0

0.2 0.4 0.6 0.8 random parameter ξ1

0.8 0.6 0.4 0.2 0 0

1

0.2 0.4 0.6 0.8 random parameter ξ

1

1

(a) Refinement of the initial mesh

(b) Refined mesh for 17 samples

Figure 2. Simplex Stochastic Collocation discretization of a two–dimensional probability space.

V.

Model validation for incompressible flows

 θt model implementation is first validated for transitional flows over adiabatic flat plate test The γ − Re cases5 that have been used to calibrate the model coefficients and that are commonly used as benchmarks for transition models. The examples study the influence of the Reynolds number, the inlet turbulence intensity, and the pressure gradient on the boundary layer transition along the plate.13 In particular, three cases were considered in this work denoted as T3A, with a zero pressure gradient boundary layer, and T3C3 and T3C4 both with a pressure gradient similar to an aft loaded turbine blade. The computational domain is modeled with H-type grids with y+ values below 0.3 in the first layer above the walls. Fig. 3 shows the results for the skin friction distribution cf and the normalized turbulent kinetic energy k/U20 in the free stream with respect to Rex for the cases T3A and T3C4. The same results are obtained compared to the results in literature,5 which validates the correct model implementation for these test cases. In order to account for the measurement uncertainty of the turbulence inlet conditions, the inlet turbulence intensity T uin and the turbulence Reynolds number ReT are considered to be uncertain with independent uniform distributions; for the test case T3C3 we assume T uin =[4.86; 5.94] and ReT =[72; 88], respectively. SC results are shown in Fig. 4(a) for the mean and the 95% uncertainty interval of the free

6 of 14 American Institute of Aeronautics and Astronautics

0.016

cf

cf 0.004

experiment deterministic simulation

0.012

0.004

0.008 0.004

0.001

0.002 0.000

k/U20

0.001 -0.002 0.000 0.0E+00

1.0E+05

2.0E+05

3.0E+05

4.0E+05

0.0014 0.0012

0.003

0.002

0.0016

0

0.0008

-0.004

0.0006 0.0004

-0.008

k/U20

0.006

0.0002 -0.012

5.0E+05

0.00E+00

5.00E+04

1.00E+05

1.50E+05

Rex

Rex (a) Case T3A with zero pressure gradient

(b) Case T3C4 with pressure gradient and separation induced transition

Figure 3. Assessment of the correct model implementation for the flat plate test cases T3A and T3C4.

stream turbulent kinetic energy k/U20 along the plate, compared to the deterministic simulation and experimental data.13 The decaying turbulent kinetic energy with Rex results in a decreasing uncertainty interval from the uncertain inlet condition for T uin . This behavior is confirmed by the decreasing standard deviation in Fig. 4(b). The deterministic result is close to the mean value and in agreement with the experiments. The accuracy of the UQ approximations of Fig. 4 is shown in Fig. 5. The quadrature error in the approximation of the mean and standard deviation for a CC rule with np points is estimated by the difference with the results for (np − 2) CC points. This implies for np = 3 that the mean is compared to the deterministic value for np = 1. The error estimate has a maximum near the start of the plate at Rex = 0 of 0.61% for the standard deviation. −3

0.1 0.09

n =3 p

n =5 p

0.08

x 10

n =5

4.5

p

std of free stream kinetic energy [%]

mean of free stream kinetic energy [%]

5

deterministic n =3 experiment

0.07 0.06 0.05 0.04 0.03 0.02 0.01

p

4 3.5 3 2.5 2 1.5 1 0.5

0 0

1

2

3 Rex

4

5

0 0

5

1

2

x 10

(a) Mean and 95% interval

3 Rex

4

5

5

x 10

(b) Standard deviation

Figure 4. Effect of uncertainty on the free stream turbulent kinetic energy (case T3C3).

The uncertainty in T uin and ReT has a different effect on the skin friction cf as function of Rex of Fig. 6. The mean and the 95% interval in Fig. 6(a) show that T uin and ReT have a small effect on the transition location. The standard deviation has a maximum in Fig. 6(b) in the transition region. The errors are also largest in the transition region. The accuracy of the mean approximation improves significantly from np = 3 to np = 5. The error in the standard deviation is with a maximum of 10.2% at the transition point also of an acceptable magnitude.

VI.

Boundary layer transition on a transonic turbine guide vane

The turbine guide vane considered in this work has been experimentally investigated by Arts et al.1 in order to study the influence of Mach number, turbulence intensity, and Reynolds number on the transitional heat transfer distribution h = q˙w /(T∞ − Tw ). The geometry of the blade is defined by the chord=67.647mm, pitch to chord ratio=0.85, throat to chord ratio=0.2207, and a stagger angle of 55 deg measured from the axial direction. The total inlet temperature is set at T0 =420K and the wall temperature is considered to 7 of 14 American Institute of Aeronautics and Astronautics

−4

2

x 10

mean (n =3) p

mean (n =5)

error in free stream kinetic energy

p

std (n =5) p

1

0 0

1

2

3 Rex

4

5

5

x 10

Figure 5. Quadrature error in the mean and standard deviation approximation for the free stream turbulent kinetic energy using SC.

−4

0.01 0.009

p

std of skin friction cf

0.006 0.005 0.004 0.003

3 2.5 2 1.5 1

0.001

0.5 3 Rex

4

5

f

p

3.5

0.002

2

n =5

4

experiment

1

n =3 p

n =5

0.007

0

x 10

4.5

p

0.008

mean of skin friction c

5

deterministic n =3

0

1

5

2

x 10

(a) Mean and 95% interval

3 Rex

4

5

5

x 10

(b) Standard deviation

Figure 6. Effect of uncertainty on the skin friction coefficient cf (case T3C3).

−4

1.5

x 10

mean (np=3) mean (n =5) p

std (np=5)

error in cf

1

0.5

0

1

2

3 Rex

4

5

5

x 10

Figure 7. Quadrature error in the mean and standard deviation approximation for the skin friction coefficient cf using SC.

8 of 14 American Institute of Aeronautics and Astronautics

be close to a constant value of 300K in the experiments. The flow is transonic with an exit Mach number close to one. The incoming turbulence level T uin was measured 55mm upstream of the leading edge plane. The uncertainty on the heat transfer measurement was of the order of ±5%. As no turbulent length scale or dissipation was measured the turbulence Reynolds number ReT is considered to be uncertain within a reasonable range, estimated using the correlation of Roach.11 The computational domain consists of 25,000 control volumes and the y + values are less than one for all cells at the wall surface. VI.A.

Initial accuracy assessment of SC for the MUR241 test case

Preliminary SC results, primarily to assess the accuracy of this collocation method, are given for the test case MUR241 with a Reynolds number (based on the chord length and exit conditions) of Re = 2.1e6 and T uin = 6% and MUR235 with Re = 1.1e6 and T uin = 6%. The stochastic results for the surface heat transfer are given in Fig. 8 for the MUR241 case in comparison with experimental data1 as function of the curvilinear coordinate along the blade s/c normalized by the chord. The reported 5% uncertainty in the experimental results is indicated by the uncertainty bars. Positive values of s/c denote the suction side of the airfoil and the pressure side is parametrized by negative s/c-values. The experiment indicates transition on the suction side at s/c = 0.6, while on the pressure side no clear transition location can be identified due to the smooth increase of the heat transfer towards the trailing edge. The uncertainty in T uin and ReT in the ranges [5.7; 6.3] and [50; 150] has a significant effect on the transition location on the suction side in Fig. 8(a). Consequently the 95% uncertainty interval in the transition region at s/c = 0.6 shows a large range of [130.0; 1370.2], which increases with np . The same divergence of the standard deviation with increasing np at s/c = 0.6 can be seen in Fig. 8(b). The large overshoot in computed heat transfer at the rear part of the suction side needs to be accounted for by model form uncertainty in addition to the varying turbulence inflow conditions. 1500

500 450 400

1000

std of heat transfer

mean of heat transfer

1250

750

500 deterministic n =3

250

0

0.2 s/c

0.4

0.6

p

350 300 250 200 150

50

p

−0.8 −0.6 −0.4 −0.2

n =5

100

p

n =5 0

deterministic np=3

experiment 0.8 1 1.2

0

(a) Mean and 95% interval

−0.8 −0.6 −0.4 −0.2

0

0.2 s/c

0.4

0.6

0.8

1

1.2

(b) Standard deviation

Figure 8. Heat transfer for the VKI test case MUR241.

When the 95% uncertainty interval is compared to the CC samples in Fig. 9, it is observed that the interval exceeds all the samples at s/c = 0.6. This is not to be expected since the CC points include the extreme values of the uncertain input ranges [5.7; 6.3] and [50; 150]. In Fig. 10 it is shown that these diverging overshoots are caused by the oscillatory global polynomial SC interpolation of the CC points of the discontinuous response surface. These oscillations increase with np such that the results do not convergence for np = 5 as shown for the cumulative distribution function (CDF) at s/c = 0.6 in Fig. 11(a). This leads to an 18.1% error in the local standard deviation at s/c = 0.6 for np = 5 in Fig. 11(b). It can also be concluded from Fig. 10 that the uncertainty in the turbulence intensity T uin has no significant effect on the heat transfer. In the following, the aleatoric uncertainty in ReT is, therefore, combined with epistemic uncertainty in the laminar-turbulent transition model, while the inlet turbulence intensity T uin is kept constant using its nominal value.

9 of 14 American Institute of Aeronautics and Astronautics

1500

95% interval of heat transfer

1250

1000

750

500 n =3 p

250

0

n =5 p

−0.8 −0.6 −0.4 −0.2

0

0.2 s/c

0.4

0.6

samples experiment 1 1.2

0.8

1500

1500

1250

1250

1000

1000

heat tranfer

heat tranfer

Figure 9. Samples and 95% interval of the heat transfer for the VKI test case MUR241.

750 500 250

750 500 250

0 150

0 150 125 100 75 50

Re

T

5.7

5.8

6.1

6

5.9

6.2

6.3

125 100 75 50

Re

Tu

T

(a) np = 3

5.7

5.9

5.8

6.2

6.1

6

6.3

Tu

(b) np = 5

Figure 10. Samples and response surface approximation of the heat transfer for the VKI blade at s/c = 0.6.

200

1

160

0.8 0.7

error in heat transfer

cumulative probability distribution

0.9

0.6 0.5 0.4 0.3 0.2 0.1 0 250

500

750 heat transfer

1000

p

100 80 60 40 20

(a) Cumulative distribution function (CDF)

std (n =5)

120

n =5 95% percentiles 1250 1500

mean (np=5)

140

np=3 p

0

mean (n =3) p

180

0

−0.8 −0.6 −0.4 −0.2

0

0.2 s/c

0.4

0.6

0.8

1

1.2

(b) Quadrature error in the mean and standard deviation approximation

Figure 11. Cumulative distribution function (CDF) and quadrature error in the mean and standard deviation of the heat transfer for the VKI test case MUR241.

10 of 14 American Institute of Aeronautics and Astronautics

VI.B.

Initial accuracy assessment of SC for the MUR235 test case

For the MUR235 case of the VKI turbine guide vane1 the turbulence Reynolds number ReT is uniformly distributed in [600; 900] and uncertainty is injected in Eq. (10) that determines the onset of transition. No transition is observed in the experiment along the pressure side, while a sharp increase of the heat transfer at s/c = 0.8 clearly indicates transition at the suction side. The epistemic uncertainty in the transition model accounts to a large extent for the difference between the deterministic simulation and the experiments in Fig. 12. No uncertainty has been added for the spot growth parameter in this preliminary result. It is clearly visible in Fig. 12(b) that the uncertain transition model injects standard deviation in the transition region. This also significantly affects the prediction of the mean heat transfer at s/c = 0.6 in the transition region, from the deterministic value of 298.4 to 373.6 and 417.8 for np = 3 and np = 5. The increasing mean value with respect to the deterministic value of 298.4 can also be deducted from the approximation of the CDF in Fig. 13(a). The error estimate for the standard deviation approximation with np = 5 indicates a relative error of 12.7% at s/c = 0.6, see Fig. 13(b). 1500

500

deterministic n =3

450

p

n =5

1250

p

400

std of heat transfer

mean of heat transfer

experiment 1000

750

500

deterministic n =3 p

np=5

350 300 250 200 150 100

250

50 0

−0.8 −0.6 −0.4 −0.2

0

0.2 s/c

0.4

0.6

0.8

1

0

1.2

−0.8 −0.6 −0.4 −0.2

(a) Mean and 95% interval

0

0.2 s/c

0.4

0.6

0.8

1

1.2

0.8

1

1.2

(b) Standard deviation

Figure 12. Heat transfer for the VKI test case MUR235.

200

1

0.7

error in heat transfer

cumulative probability distribution

160

0.8

0.6 0.5 0.4 0.3 0.2 0.1 0 298

500

750 heat transfer

1000

p

100 80 60 40 20

(a) Cumulative distribution function (CDF)

std (n =5)

120

n =5 95% percentiles 1250 1500

mean (np=5)

140

np=3 p

0

mean (n =3) p

180

0.9

0

−0.8 −0.6 −0.4 −0.2

0

0.2 s/c

0.4

0.6

(b) Quadrature error in the mean and standard deviation

Figure 13. Cumulative distribution function (CDF) and quadrature error in the mean and standard deviation of the heat transfer for the VKI test case MUR235.

VII.

Results for the VKI MUR 235 and MUR 241 cases using SC and SSC

Because of the diverging results for Stochastic Collocation (SC) in the transition region the convergence is computed up to more quadrature points and the results are compared to those of Simplex Stochastic Collocation (SSC) hereafter. Since the epistemic uncertainty in the transition model had significant impact on the results for the transonic turbine guide vane, also a second source of epistemic uncertainty is considered

11 of 14 American Institute of Aeronautics and Astronautics

in terms of the turbulent spot growth parameter as described in section III. Therefore, three uncertainties are considered: the inlet turbulent length-scale given by ReT , the correlation that determines the transition onset location and the correlation for the turbulent spot growth within the transitional boundary layer. VII.A.

MUR241 test case

In this case the turbulence Reynolds number ReT is uniformly distributed in [140; 260]. For SC, three levels of Clenshaw–Curtis rules are considered with np = {3, 5, 9}, which leads for three uncertainties to n = {27, 125, 729} samples. The results of SSC are shown for n = {10, 50, 100} samples. Fig. 14 shows that the turbulent spot growth parameter has a significant impact on the heat transfer downstream of the transition point on the suction side of the blade in terms of the mean and the 99% interval. The epistemic uncertainty in the turbulent spot growth modeling, therefore, gives a good explanation for the discrepancy between the simulation and the experiment downstream of the transition point in Fig. 8. It leads for SC in overshoots in the transition region and results that do not show convergence for up to n = 729 samples. The results of SSC are better behaved with good agreement between the prediction with 50 and 100 samples. Similar observations can be made from the comparison of the results for the standard deviation in Fig. 15. 1500

1500

deterministic np=3 (n=27) n =5 (n=125)

1250

deterministic n=10 n=50 n=100 experiment

1250

p

n =9 (n=729) experiment

1000

mean of heat transfer

mean of heat transfer

p

750

500

1000

250

0

750

500

250

−0.8 −0.6 −0.4 −0.2

0

0.2 s/c

0.4

0.6

0.8

1

0

1.2

(a) Stochastic Collocation

−0.8 −0.6 −0.4 −0.2

0

0.2 s/c

0.4

0.6

0.8

1

1.2

1

1.2

(b) Simplex Stochastic Collocation

Figure 14. Mean and 95% interval of the heat transfer for the VKI test case MUR241.

500

np=3 (n=27)

450

n =5 (n=125)

450

400

np=9 (n=729)

400

p

350

std of heat transfer

std of heat transfer

500

300 250 200 150

350 300 250 200 150

100

100

50

50

0

−0.8 −0.6 −0.4 −0.2

0

0.2 s/c

0.4

0.6

0.8

1

0

1.2

(a) Stochastic Collocation

n=10 n=50 n=100

−0.8 −0.6 −0.4 −0.2

0

0.2 s/c

0.4

0.6

0.8

(b) Simplex Stochastic Collocation

Figure 15. Standard deviation of the heat transfer for the VKI test case MUR241.

VII.B.

MUR235 test case

For the case MUR235 ReT has been estimated to be uniformly distributed between [490; 910]. The additional epistemic uncertainty in the turbulent spot growth has a smaller effect on the results in Fig. 16 for ReT in 12 of 14 American Institute of Aeronautics and Astronautics

[490; 910] compared to MUR241 presented above. However, it results in a larger spread of the SC predictions, while SSC shows converged results at n = 100 for the mean and the uncertainty interval. The standard deviation of the heat transfer in Fig. 17 confirms these observations. 1500

1500

deterministic n =3 (n=27)

deterministic n=10 n=50 n=100 experiment

p

n =5 (n=125)

1250

1250

p

n =9 (n=729) experiment

1000

750

500

mean of heat transfer

mean of heat transfer

p

1000

250

0

750

500

250

−0.8 −0.6 −0.4 −0.2

0

0.2 s/c

0.4

0.6

0.8

1

0

1.2

(a) Stochastic Collocation

−0.8 −0.6 −0.4 −0.2

0

0.2 s/c

0.4

0.6

0.8

1

1.2

1

1.2

(b) Simplex Stochastic Collocation

Figure 16. Mean and 95% interval of the heat transfer for the VKI test case MUR235.

500

500

n =3 (n=27) p

450

450

p

n =9 (n=729) p

300 250 200 150

350 300 250 200 150

100

100

50

50

0

−0.8 −0.6 −0.4 −0.2

n=10 n=50 n=100

400

350

std of heat transfer

std of heat transfer

400

n =5 (n=125)

0

0.2 s/c

0.4

0.6

0.8

1

0

1.2

(a) Stochastic Collocation

−0.8 −0.6 −0.4 −0.2

0

0.2 s/c

0.4

0.6

0.8

(b) Simplex Stochastic Collocation

Figure 17. Standard deviation of the heat transfer for the VKI test case MUR235.

VIII.

Conclusions

The ability to accurately predict the transition process is crucial for the design of efficient and reliable transonic gas turbine compressors. The boundary layer development, losses, efficiency, and heat transfer are known to be sensitive to small physical variations and uncertainties in the transition model. In this work, the impact of these epistemic uncertainties as well as the aleatory uncertainties is, therefore, quantified in the numerical uncertainty analysis of a transonic turbine guide vane. The transition model implementation is first validated for the skin friction distribution cf and the normalized turbulent kinetic energy k/U02 in the free stream in the transitional flow over adiabatic flat plate benchmarks T3A and T3C4. The uncertainty analysis results for cf show that uncertainty in the inlet turbulence intensity T uin and the turbulence Reynolds number ReT have a small effect on the location of the transition point. In the turbine guide vane test case MUR241, uncertainty in ReT has a significant effect on the transition location on the suction side. The impact of the variation in the turbulence intensity T uin is negligible for this case. The aleatoric uncertainty in ReT and the epistemic uncertainty in the laminarturbulent transition model for the MUR235 case shows that the model uncertainty accounts to a large extent for the difference between the deterministic simulation and the experiments. The uncertain transition model injects uncertainty in the transition region with a significant effect on the mean and standard deviation of

13 of 14 American Institute of Aeronautics and Astronautics

the surface heat transfer. The additional epistemic uncertainty in the transonic spot growth parameter explains the discrepancy between the simulations and the experiments in the heat transfer downstream of the transition on the suction side of the blade for the MUR241 case. This problem with three uncertainties in ReT (inlet turbulent lengthscale) and epistemic parameters UReθt (transition onset correlation) and UP (turbulent spot growth) gives overshoots and results that are not yet converged for Stochastic Collocation with n = 729 samples. The comparison with Simplex Stochastic Collocation demonstrates the more robust convergence for n = 100 samples of the latter method.

Acknowledgments This material is based upon work supported by the Department of Energy (National Nuclear Security Administration) under Award Number NA28614.

References 1 T.

Arts, M. Lambert de Rouvriot, A.W. Rutherford, Aero-Thermal Investigation of a Highly Loaded Transonic Linear Turbine Guide Vane Cascade, Technical Note, 174, von Karman Institute for Fluid Dynamics, Belgium, 1990. 2 I. Babuˇ ska, F. Nobile, R. Tempone, A stochastic collocation method for elliptic partial differential equations with random input data, SIAM J. Numer. Anal. 45 (2007) 1005–1034. 3 P. Batten, M.A. Leschziner, U.C. Goldberg, Average-state Jacobeans and implicit methods for compressible viscous and turbulent flows, J. Comput. Phys. 137 (1997) 38–78. 4 N.P. Gopinathrao, C. Mabilat, S. Alizadeh, Non-deterministic thermo-fluid analysis of a compressor rotor-stator cavity, 50th AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference, Palm Springs, California (2009) AIAA-2009-2278. 5 R.B. Langtry, F.R. Menter, Correlation-based transition modeling for unstructured parallelized computational fluid dynamics codes, AIAA J. 47 (2009) 2894–2906. 6 G.J.A. Loeven, H. Bijl, The application of the probabilistic collocation method to a transonic axial flow compressor, 51th AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference, Orlando, Florida (2010) AIAA-2010-2923. 7 F.R. Menter Two-equation eddy-viscosity turbulence models for engineering applications, AIAA J. 32 (1994) 1598–1605. 8 R. Peˇ cnik, W. Sanz, Application of the turbulent potential model to heat transfer predictions on a turbine guide vane, J. Turbomach. 129 (2007) 628–635. 9 R. Peˇ cnik, P. Constantine, F. Ham, G. Iaccarino, A probabilistic framework for high-speed flow simulations, Annual Research Briefs, Center for Turbulence Research, Stanford (2008) 3–17. 10 R. Peˇ cnik, V.E. Terrapon, F. Ham, G. Iaccarino, Full system scramjet simulation, Annual Research Briefs, Center for Turbulence Research, Stanford (2009) 33–45. 11 P.E. Roach, The generation of nearly isotropic turbulence by means of grids, Heat Fluid Flow. 8 (1989) 82–92. 12 B. Satish, K. Buschelman, V. Eijkhout, W.D. Gropp, D. Kaushik, M.G. Knepley, L.C. McInnes, B.F. Smith, H. Zhang, PETSc (2009) http://www.mcs.anl.gov/petsc. 13 A.M. Savill, A synthesis of T3 test case predictions, numerical simulation of unsteady flows and transition to turbulence, O. Pironneau et al. (Eds.), Cambridge University Press, Cambridge, 404–442. 14 J. Steelant, E. Dick, Modeling of laminar-turbulent transition for high freestream turbulence, J. Turbomach. 123 (2001) 22–30. 15 E.F. Toro, M. Spruce, W. Spears, Restoration of the contact surface in the HLL-Riemann solver, Shock Waves 4 (1994) 25–34. 16 J.A.S. Witteveen, G. Iaccarino, Simplex elements stochastic collocation for uncertainty propagation in robust design optimization, 48th AIAA Aerospace Sciences Meeting, Orlando, Florida (2010) AIAA-2010-1313. 17 J.A.S. Witteveen, G. Iaccarino, Simplex elements stochastic collocation in higher-dimensional probability spaces, 51st AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference, Orlando, Florida (2010) AIAA2010-2924. 18 D. Xiu, J.S. Hesthaven, High-order collocation methods for differential equations with random inputs, SIAM J. Sci. Comput. 27 (2005) 1118–1139.

14 of 14 American Institute of Aeronautics and Astronautics

Uncertainty Quantification for Laminar-Turbulent ...

entirely on empirical correlations obtained from existing data sets for simple flow configurations. ... Postdoctoral Fellow, Center for Turbulence Research, Building 500, Stanford University, Stanford, CA ... 4 - 7 January 2011, Orlando, Florida.

550KB Sizes 0 Downloads 267 Views

Recommend Documents

Uncertainty Quantification for Stochastic Subspace ...
Uncertainty Quantification for Stochastic Subspace Identification on. Multi-Setup Measurements. Michael Döhler, Xuan-Binh Lam and Laurent Mevel. Abstract— ...

Uncertainty Quantification for Laminar-Turbulent ... - Jeroen Witteveen
Jan 7, 2011 - 8. Considerable effort has been spent in the past two decades to develop transition models for engineering applications to predict transitional ...

Uncertainty Quantification for Multi-Frequency ...
reliable predictions, which can be utilized in robust design optimization and reducing .... on a 60 × 15 × 30m domain using an unstructured hexahedral mesh. ... Results for the time evolution of the mean µL(t) and the standard deviation σL(t) of 

Uncertainty Quantification for Stochastic Damage ...
finite element model of the structure. Damage localization using both finite element information and modal parameters estimated from ambient vibration data collected from sensors is possible by the Stochastic Dynamic Damage Location Vector (SDDLV) ap

efficient uncertainty quantification in unsteady ...
of the probability distribution and statistical moments µui (x,t) of the output u(x, t, a), .... subject to a symmetric unimodal beta distribution with β1 = β2 = 2 with a ...

Efficient Uncertainty Quantification in Computational Fluid-Structure ...
Sep 21, 2007 - Abstract. Uncertainty quantification in complex flow and fluid-structure interaction simulations requires efficient uncertainty quantification meth-.

Uncertainty Quantification for the Trailing-Edge Noise of ...
on the restricted domain with the above extracted velocity profiles, directly yields .... from LWT RANS computations and (dash-dot) uncertainty bounds around inlet ..... G., Wang, M. & Roger, M. 2003 Analysis of flow conditions in free-jet experi-.

Quantification of uncertainty in nonlinear soil models at ...
Recent earth- quakes in Japan ... al Research Institute for Earth Science and Disaster. Prevention ..... not predicting a large degree of nonlinear soil be- havior.

Uncertainty quantification and error estimation in ... - Semantic Scholar
non-reactive simulations, Annual Research Briefs, Center for Turbulence Research, Stanford University (2010) 57–68. 9M. Smart, N. Hass, A. Paull, Flight data ...

a robust and efficient uncertainty quantification method ...
∗e-mail: [email protected], web page: http://www.jeroenwitteveen.com. †e-mail: ... Numerical errors in multi-physics simulations start to reach acceptable ...

Uncertainty quantification and error estimation in scramjet simulation
is overwhelmed by abundant uncertainties and errors. This limited accuracy of the numerical prediction of in-flight performance seriously hampers scramjet design. The objective of the Predictive Science Academic Alliance Program (PSAAP) at Stanford U

Uncertainty Quantification in MD simulations of ...
that, based on the target application, only certain molecules or ions can pass ...... accordance with Figure 8 (b) which showed that the noise level in the Cl− ..... Figure 12 shows that the predictive samples form a “cloud” demonstrating that

Uncertainty quantification and error estimation in scramjet simulation
V.A. Aleatoric flight conditions. The flow conditions of the HyShot II flight are uncertain due to the failure of the radar tracking system during the experiment. Therefore, the free stream conditions were inferred from pressure measurements in the u

Quantification of uncertainty in nonlinear soil models at ...
Jun 17, 2013 - DEEPSOIL. – ABAQUS. • Equivalent-linear models: Within SHAKE, the following modulus-reduction and damping relationships are tested: – Zhang et al. (2005). – Darendeli (2001). • Nonlinear models: - DEEPSOIL (Hashash et al., 20

Uncertainty quantification in MD simulations of ...
the accuracy with which the potential function can reproduce the atomic .... σ) and Coulombic charge, in multiples of the electron charge |e|, for each atom type.

Uncertainty Quantification in Fluid-Structure Interaction ...
the probability distribution and statistical moments µui (x,t) of the output u(x, t, a), which ...... beta distribution with a coefficient of variation of 1% around a mean of ...

Uncertainty quantification and error estimation in scramjet simulation
to improve the current predictive simulation capabilities for scramjet engine flows. ..... solution is obtained by the automatic differentiation software package ...

Unary quantification redux
Qx(Ax) is true iff more than half of the entities in the domain of quantification ... To get a feel for the distinctive features of Belnap's system, we present a simplified.

Unary quantification redux
Qx(Ax) is true iff more than half of the entities in the domain of quantification ... formula as asserting the proposition with the free variables x referring to the ... strate its correctness, let us first check some examples to see how Belnap's ana

CONTROLLING UNCERTAINTY EFFECTS Uncertainty ...
Conflicts from Ireland to Afghanistan are a tragic testament to the power of religious attitudes, and we ..... a short survey as part of a psychology class project. They first ... The values options included: Business/ Economics/ Making Money, ...

NUMERICAL METHODS FOR UNCERTAINTY ...
A straightforward application of Monte Carlo (MC) method leads to the .... obviously not the case in the PC approach that requires the development of a modified.