3.1. Mesh Convergence Study
First, a mesh convergence study for LES at
was performed to determine a mesh that satisfactorily balances accuracy and computing resources. The underlying mesh refinement strategy is shown in
Figure 5c. The mesh is linearly refined towards the wall with a constant factor. The normal wall distance of the first cells
and
, which correspond to the dimensionless wall distances
and
, are kept constant. This ensures correct resolution of the wall-boundary layer. The wall tangential cell dimensions
and
at the central planes of the cavity are gradually refined based on values for the dimensionless tangential cell size
and
from best practice guidelines that can be found in the literature [
31]. The recommendations are
,
, and
for the wall-normal, streamwise, and spanwise grid resolutions. Through the stepwise refinements of
and
through
and
, the required number of cells in the boundary layer and the required wall tangential resolution can be determined, which are crucial for correct calculation. It is possible to vary the two parameters independently of each other in order to determine the appropriate wall tangential and wall normal resolution. To keep the computational effort reasonable, the parameters were changed simultaneously, assuming that the required dimensionless mesh resolution is identical on the enclosing walls. An appropriate mesh for the large parametric analysis balancing accuracy and computational effort was finally determined by using the local Nusselt number as a convergence criterion:
. For this purpose, the time-averaged Nusselt number profile and the local standard deviation were investigated. The results were obtained by averaging the profiles on the left and right walls and by averaging in the
z-direction for the 3D LES simulation. For time averaging, the time
[
13] was used. Afterwards, the LES results were compared with the DNS results using the mean absolute error
[
32,
33]. The DNS was defined as a reference. avg and sdev denote the average value and standard deviation, respectively.
denotes the temporal and spatial average in the periodic direction.
is the fluctuating component according to
with a flow property
and its temporal mean component
. From this, the evaluation is described by the following expressions:
Figure 6a–d clearly shows that, with increasing refinement, the LES results are in very good agreement with the DNS results.
Figure 6a,b shows the normalized MAE of the LES from the DNS results regarding the average and standard deviation of
, which gradually decreases when the mesh is refined. It can be seen in
Figure 6a,b that, for
, the difference is very close to zero. The time-averaged local
number profile in
Figure 6c already shows good results with a coarse resolution of the domain. There are minor differences near the top and bottom of the cavity. A detail plot within
Figure 6c shows the gradual convergence of the LES to the DNS profile. However, for a coarse grid, the transition point, which corresponds to the point with maximum fluctuations and thus the largest standard deviation, is not predicted properly, as can be seen in
Figure 6d. In addition, no recirculation zone is formed. The formation of the recirculation zone, as previously described by Trias et al. [
13], only takes place in the case of fine grids.
Due to the fact that the recirculation zone does not develop and the transition occurs further downstream for coarse grids, the fluid is diverted by the horizontal walls instead and causes fluctuations when it hits the vertical walls.
To define a mesh that is accurate enough to reflect the global characteristics compared to the DNS, an upper limit for the relative deviation between LES and DNS was defined:
with
[
13].
denotes the spatial average in the
y-direction. This condition is fulfilled from a grid resolution of
. Therefore, this mesh resolution was selected for the following investigations. The same spatial resolution was chosen for the LES and URANS to ensure that the spatial discretization error is comparable and provides a direct comparison of the methods independent of the discretization techniques. When considering the LES results, the structure and spatial resolution of the boundary layer were investigated, as can be seen in
Figure 7. The thickness of the velocity boundary layer
was determined by the occurring inflection point
in the velocity profile, which is characteristic for natural convection. The temperature boundary layer
was defined by a gradient criterion
:
. These criteria made it possible to investigate the velocity and temperature boundary layer thickness over the entire height in a consistent manner and to determine the required number of cells in the boundary layer. The maximum root-mean-square (rms) of the velocity, which takes place after reaching the maximum velocity, is approximately in the area of the maximum velocity gradient, characterized by the inflection point. Therefore, the definition by the inflection point and the rms criterion are in good agreement at the middle height plane, as shown in
Figure 7a. The situation is analogous for the temperature boundary layer. To achieve comparable results to the rms criterion, the gradient criterion can be adjusted accordingly. From the results the mean cell number was determined with the following expression:
, where
is the vertical cell length of the corresponding cell,
is the respective number of cells in the boundary layer, and
is the mean number of cells in the boundary layer over the entire height
H. With the previously defined criteria, this results in a velocity and temperature boundary layer resolution of
and
respectively. The determination of the required wall normal resolution with
and
and the required wall tangential resolution
and
with the present mesh refinement strategy enables the transfer of results to problems with similar physics such as the THAI containment.
3.2. Quantities of Interest
Before the assessment of uncertainties and analyzing sensitivities, the result quantities, which adequately describe the underlying physics of the mixing process in the cavity, have to be defined. In the context of uncertainty quantification, these are referred to as Quantities of Interest (QoI). For this purpose, integral scalar quantities provide a plain description of the transient profiles of the mixing process. In this way, summarizing statements can be made about the complete transient. Since the flow is driven by buoyancy effects, the spatial averaged Nusselt number
over the respective walls is of paramount importance. In this way, the convective heat transfer in the cavity is evaluated and the transients during the mixing process can be examined more closely. It is determined by the expression
where
denotes the wall-normal unit vector. The effects on the convection mechanisms are described by global kinetic energy.
is the quotient of the global kinetic energy by a reference kinetic energy
, which contains the material properties of air:
m denotes the mass and
M denotes the total mass of the fluid domain. Due to the thermal boundary conditions, natural convection develops within the cavity and the helium stratification is eroded. This mixing process continues until a homogeneous mixture finally forms. The achievement of this state can be quantified by the mixture uniformity
, which is the volume-weighted standard deviation of the mole fraction
X from the homogeneous equilibrium state mole fraction
over the whole fluid domain. The initial mixture uniformity
is the highest occurring standard deviation during the mixing process. When varying the initial mole fraction difference
, the initial mixture uniformity
changes. Therefore, for normalization, the initial mixture uniformity
of the reference case, which corresponds to the initial conditions of a stratification with constant mole fraction, is used. Following the definition of Danckwerts [
34], this gives the expression for the segregation intensity, described by the Equation (
12):
From the segregation intensity
I, the definition of the mixing intensity
[
35] can also be derived for the interpretation of the results. In the reference case,
describes a completely inhomogeneous mixture and
characterizes a completely homogeneous mixture and vice versa for the mixing intensity
. Since the initial state changes, when the mole fraction difference
is varied, the initial segregation intensity
I is also different and the mixture transients are consistently captured with all occuring changes. Finally, a criterion for the mixing time can be derived from
I. A completely homogeneous mixture is characterized by
, and therefore by definition of an upper bound
, which
I has to fall below, the achievement of this state can be quantified. When considering mass transfer processes, the Fourier number
enables a dimensionless description of time, where
is the diffusion coefficient of the reference case. Hence, the time, when the homogeneous state is reached, can be described by
. The time-dependent quantities described above are combined to the result variables vector
, where the bold notation indicates a vector or matrix.
Together with
, additional scalar quantities that describe the mixing process were derived. For this purpose, the integral mean value
of the described quantities
(Equation (
13)) is formed over the respective mixing time in Equation (
14).
Next to the actual integral mean value over the mixing process, this also provides a measure of the shape of the mixing transient because transients, which can be mapped on each other by linear stretching or compression of the time coordinate, have an identical integral mean value. Therefore, occurring deviations also indicate a change in the shape of the mixing transient. Because different transients can have the same integral mean value
, the mean absolute deviation
from the reference case over the reference mixing time
was therefore defined in Equation (
15).
All differences in the mixing transient are captured over the mixing time of the reference case and provide a good measure for the variability. At the same time, it can be determined how big the difference between the adiabatic configuration and a configuration with temperature specification is. Finally, all variables under investigation are summarized in a matrix in Equation (
16).
The results for the reference case are different for LES and URANS. The reference values
are applied to the LES results and the reference values
are applied to the URANS results. As listed in Equation (
17), the mean integral values of the reference case
are applied for the normalization of the later results, since the mean absolute deviation
from the reference case itself is zero.
3.3. Sensitivity Analysis
For the investigation of uncertainties, identification of the most influential parameters is of great importance, since uncertainties occurring in these parameters mainly dominate the uncertainty in the results. Variance-based methods also enable quantification of the proportion of the result uncertainty caused by individual parameters. Therefore, uncertainty analysis is closely related to sensitivity analysis. For assessment of the sensitivities, the Morris method is applied [
36].
denotes the parameter vector with the elements
. The defined parameter space of the computational model is explored by
r-trajectories, for which the initial points are gained by random sampling, and then one parameter after the other is changed step by step. The respective trajectory is described by the index
j. Then, for each trajectory, the elementary effects
are calculated, which represent the gradient when varying the
ith parameter using the forward difference. From this, the mean and modified mean of the elementary effects
and
as well as the standard deviation of the elementary effects
are determined. The modified mean indicates the overall effect and the standard deviation indicates nonlinear or interaction effects between the input parameters. For evaluation of the sensitivity, both parameters are important because a low value of
combined with a considerable value of
means that there is a great impact in certain regions of the parameter space and that the corresponding parameter should be taken into account for the subsequent analysis. The expressions for the calculation [
36] are
where
is the
ith unit vector and the parameter step is
with the input space uniformly partitioned into the even number of levels
p.
denotes the parameter interval width vector and
describes the scalar parameter step. Global sensitivity analysis requires a large number of trajectories
r per parameter. However, a large number of parameters in the mixing process are involved, which might be subject to uncertainties. Therefore, the total number of parameters was divided into three parameter vectors to maintain clarity for the further steps, such that the vector of input parameters is given by
with:
First, the most important parameters are extracted by preselection with a one-at-a-time (OAT) approach. Starting from an initial point
in parameter space, the parameters were changed along each orthogonal dimension with two trajectories
, respectively. The initial point was defined as follows:
takes the initial values from
Section 2.2 and according to
Table A1. The most influential parameters were determined by assuming an identical parameter interval, which has the width of 20 percent with respect to the corresponding nominal value in each parameter. The associated values for
and
result according to
Table 1. Through this definition, the parameters are varied symmetrically around the defined initial point with the assumed standard deviation
in the parameters according to
Section 3.4. The standard deviation is representative for the total dispersion around the mean, and in this way, an estimation for the mean impact of single parameters is achieved.
From the results, the modified mean
can be obtained. The standard deviation
was not taken into account here, as the number of trajectories is small. Here, the main criterion for the evaluation of the sensitivity was the mixing time measured by the dimensionless Fourier number
. It can be seen in
Figure 8 that the material values such as the heat capacity
and
, the dynamic viscosities
and
, as well as
have no great influence on the mixing time for
. Additionally, the model coefficients of the gradient flux approach
and
do not show major impacts on the mixing time. On the other hand,
causes a medium impact, since the main proportion of the mixture is air. However, it is assumed that there is a small uncertainty in the prediction of material values. The thermal boundary conditions via
,
, and
show great influence on the mixing time, as the mixing process is primarily based on buoyancy effects. The diffusion coefficient via
also leads to a large deviation. The change in the helium stratification through
also causes relevant differences in the mixing time.
On the basis of these findings, the five most important parameters were selected. These are investigated next by Global Sensitivity Analysis (GSA) using the Morris method according to Equation (
18). Consequently, the parameter vector, which contains the prioritized parameters, gives the following expression:
The input space is uniformly partitioned into four levels
. For the parameter intervals in
Table 2 scaled to
, this results in the scalar parameter step
. A random initial point
is chosen from the set
for each parameter. Starting from this point, the parameters are changed in an arbitrary order. The assessment of the sensitivity was conducted by
trajectories. In
Table 2, the intervals under investigation are listed. The parameter intervals were defined according to the probability density functions according to Equation (
31) in
Section 3.4. With
and
denoting the expectation and standard deviation of the random input variables
, the interval was defined by
as it contains the majority of all possible parameter values. The aim of this study was to capture local interaction effects and possible nonlinearities using a reasonable number of trajectories. With a constant number of levels
p, a wide interval width would lead to larger parameter steps, which cause less accurate prediction of local effects. For this reason, the parameter intervals were chosen in this way.
All previously described result variables according to
Section 3.2 are examined. The integral mean Nusselt number
measures the convective heat transfer between the fluid and the enclosing walls during the mixing process, which significantly influences the mixing process. Heat is supplied on the left wall and dissipated on the right wall. Cold gas coming from the right wall is heated by the lower wall and hot gas coming from the left wall is cooled by the upper wall.
and
, which are shown in
Figure 9a,b, exhibit the biggest change through the temperature specification of the associated walls via
because this simultaneously causes a change in the wall adjacent temperature gradient. The remaining parameters show medium impact, which is comparable for these parameters, when the values of the respective opposite walls are considered averaged with each other. LES and URANS show almost linear behavior regarding the integral mean Nusselt number in Equation (
14), as can be seen in
Figure 9e,f. Both approaches show good agreement. Additionally, for
and
, which are shown in
Figure 9c,d, the corresponding temperature specification via
causes the greatest influence. The influence of
is greater at the top wall than at the bottom wall because the helium is initially placed at the top and thus causes an influence during the mixing process.
leads to the smallest impact for this case. In comparison to the left and right walls,
has larger values for the bottom and top walls. There is good correspondence between LES and URANS for both
and
at the bottom and top walls.
is a quantity for measuring any occurring temporal local deviation in the convective heat transfer from the reference case and thus provides information about the influence of the parameters over the whole mixing transient. The results depicted in
Figure 10 show a similar trend to that for
. Therefore, the conclusions from
can be transferred to the results for
. An important difference exists for the measure
, which captures nonlinear effects. The target function
according to Equation (
15) evaluates the enclosed area between the mixing transients. Hence, nonlinear effects arise in the case of a strongly nonlinear profile of the transient if it is shifted through a shorter mixing time. In addition, longer mixing transients are cut off when the mixing time of the reference case is reached, which also results in nonlinearities. This fact explains the higher values in
. This approach was chosen because the fixed integration interval enables consistent normalization. An integration interval with the mixing time of the respective case would cause identical nonlinear effects with no consistent normalization opportunity. Defining a long integration interval that contains all mixing times has the disadvantage that steady-state values are taken into account, which are irrelevant for the mixing process and therefore falsify the results. The evaluation according to Equation (
15) has minor drawbacks and therefore offers a good basis for the evaluation.
The thermal boundary conditions via
,
, and
have great impact on the integral mean global kinetic energy during the mixing process
, as can be seen in
Figure 11a. The remaining parameters show minor impact.
also behaves almost linearly when the parameters are changed, as can be seen in
Figure 11c. In contrast to
, the other parameters also show noticeable influence for the deviation in kinetic energy from the reference case
, shown in
Figure 11b. This is due to the fact that, although the integral mean kinetic energy
remains the same by changing these parameters, the mixing time changes relative to the reference case and this leads to the measured deviation.
in
Figure 11d correlates to
, since the target function itself is nonlinear, and consequently, larger change yields greater associated nonlinearity.
The duration of the entire mixing process is measured by the mixing Fourier number
. As can be seen in
Figure 12a, the associated modified mean
is greatest for
and
.
has a comparable order of magnitude for the other parameters, but
and
have slightly higher
than
. In
Figure 12d, the nonlinearities or interaction effects are largest for
,
, and
. When
is changed, larger interaction behavior or nonlinearities occur with URANS compared to LES, as can be seen from
. In summary, LES and URANS show comparable sensitivity with regard to the mixing time.
The change in shape or rather the integral mean segregation intensity
according to
in
Figure 12b is largely determined by
because the initial distribution of the helium mole fraction varies. This is reflected in the mixing behavior and the shape of the mixing transient. The remaining parameters show medium impact, and the shape of the mixing transient is mostly retained. The conclusions from
can largely be applied to the mean absolute deviation in the segregation intensity
.
and
have the greatest influence, as they also have the greatest influence on
. Hence, the change in these parameters shifts the mixing transients regarding time and consequently increases
. The larger values for
in
Figure 12f are due to the nonlinearity effects of the target function
, which has already been discussed.
Since a large number of parameters enormously increases the required computing resources for the investigation of uncertainties, three of the remaining parameters will be examined further. The mixing Fourier number was selected as the main criterion to pick the most relevant parameters. The characteristic temperature difference via has the greatest importance for the buoyancy-driven mixing process. and have slightly higher impact than . For this reason, , , and were chosen for the uncertainty quantification. The diffusion coefficient via has also great importance for the mass transfer. For the application to a binary mixture, however, it is assumed that the diffusion coefficient can be described sufficiently accurate by a dynamic model and is subject to minor uncertainties. Therefore, this parameter is also neglected in the following considerations.
3.4. Uncertainty Quantification
Based on the preceding sensitivity analysis in
Section 3.3, the propagation of uncertainties in one initial condition, which is represented by the change in the initial mole fraction difference of the helium stratification through
, and two thermal boundary conditions, which are represented by the change in the characteristic temperature difference through
and the change in the wall tangential temperature gradient through
, were investigated.
The open-source software
Dakota 6.10 [
37] was used as the uncertainty quantification framework. For the evaluation with LES and URANS, non-intrusive Polynomial Chaos Expansions (PCE) were applied because of the high convergence rate of the stochastic results with an increasing number of simulation runs. Therefore, very accurate results can potentially be obtained even with a small number of calculations. The random input variables
are functions that map events
from the sample space
to realizations
. PCE is a spectral method in which random response functions
are described by suitable multidimensional orthogonal polynomials
as a function of the random input variables
. With a sequence
of random variables, the infinite expansion results in expression Equations (
22) and (
23) [
38,
39]:
There is a one-to-one correspondence between the PCE coefficients
and
and between the multidimensional orthogonal polynomials
and
. In Equation (
22), each additional set of nested summations designates a collection of polynomials with increasing order. Term-based indexing in Equation (
23) instead of order-based indexing simplifies the expression. Finally, a limited number of random variables
n and order truncation leads to Equation (
24) with
P summation terms. According to [
39], the orthogonal polynomials are generated numerically by using Gauss–Wigert [
40], discretized Stieltjes [
41], Chebyshev [
41], or Gramm–Schmidt [
42] approaches. The Gauss points and weights are computed by the Golub–Welsch [
43] tridiagonal eigensolution. This allows us to define arbitrary probability density functions for the input variables and eliminates the need to induce additional nonlinearity through variable transformations. The PCE coefficients
are estimated here by using spectral projection. The orthogonality property of the polynomials helps to extract each coefficient. The following expression [
38], which contains the inner product
on
with the weight
, gives the coefficients by
where
is the joint probability density (weight) function. The inner product
can be computed analytically. For solving the multidimensional integral in Equation (
25), the discrete projection, which is also termed pseudospectral, is applied. The multidimensional integral can be approximated by a tensor product of one-dimensional quadrature formulas. A one-dimensional quadrature operator with the level
, the quadrature points
, and a function
gives the following expression [
38]:
For the multivariate case
and a multi-index
with
, the full tensor product quadrature formula results in Equation (
27) [
38].
However, for evaluation of this full tensor product, a very large number of function evaluations is required. Therefore, multi-dimensional integration by the Smolyak sparse grid method [
44] according to Equation (
28) is performed, which tremendously reduces the number of quadrature points while a high accuracy is preserved. The sparse grid quadrature rule is defined by the following expression [
38,
39]:
The expression before the tensor product is a binomial coefficient, which is defined as follows:
The dimension independent maximum sparse grid level
m controls the number of function evaluations and the associated accuracy of the PCE. The PCE is built through a linear combination of separate tensor polynomial chaos expansions for each underlying tensor quadrature grid [
45]. Summation of the expansion terms is conducted with the Smolyak combinatorial coefficient in Equation (
28). This improves accuracy in the coefficient estimation and preserves the consistency of the PCE [
39].
In the present investigation, a sparse grid level
is considered and closed fully nested Clenshaw–Curtis points are applied for quadrature. Univariate and bivariate effects in
are modeled with orthogonal polynomials of highest-order
, resulting from the PCE construction according to [
45]. Therefore, together with the GSA in
Section 3.3, which shows the approximately linear or low nonlinear behavior of
, the results are assumed to be sufficiently accurate.
For the subsequent uncertainty analysis, a three-dimensional parameter space
with the parameter vector
contains the prioritized parameters in the following expression:
The initial and boundary conditions are often not exactly known and are therefore subject to uncertainties. Next to the initial conditions, the boundary conditions enable the unambiguous solution of a partial differential equation regarding space and time. However, in case of the Navier–Stokes equations, inherent uncertainties in the boundary conditions arise due to the interaction of turbulent fluctuations with the boundary. For this reason, a normal distribution for uncertainty in the characteristic temperature difference through
was defined according to the central limit theorem, which states that many independent random effects lead to normal distribution. The uncertainty of the wall tangential temperature gradient through
was defined to follow a log-normal distribution, which also in the course of the central limit theorem represents the distribution that results from the product of many positive independent random variables. The temperature gradient represents the temperature profile at the boundary that is established by the counterflow heat exchanger in the application case and since the supplied heat from the heat exchanger is always positive, the assumption of a log-normal distribution was made. In addition, the injection process creates large uncertainties in the actual build-up of helium stratification due to the turbulent flow of the gas stream through the injection nozzle. Hence, for the initial condition through
, a half normal distribution was also assumed because, if the stratification is stable, the density decreases in the vertical direction. The random variables for the realizations
,
, and
are
,
, and
. The random vector of parameters is
. The standard deviation of
and
was assumed to be
. For
, the standard deviation is defined with
since a large uncertainty in the structure of the stratification is assumed.
and
denotes a normal distribution and a log-normal distribution with expectation
and variance
.
denotes a truncated normal distribution with
a and
b as the lower and upper bounds. The corresponding probability density functions
of the uncertain input parameters are then as follows:
After the propagation of these uncertainties, the probability density functions (PDF)
, the expansion mean
, and the standard deviation
for LES and URANS were determined. In
Appendix B, the first-, second- and total-order Sobol indices are listed in
Table A3,
Table A4 and
Table A5 respectively. Thus, the uncertainties in the responses can be apportioned to the uncertainties in the input parameters.
was determined by sampling on the PCE approximation considering the input probability density functions
. The mean
, standard deviation
, and Sobol indices were computed analytically from the PCE coefficients
.
When comparing the mean values and the standard deviations in
Table 3, it is noticeable that the results for
and
largely coincide for LES and URANS. With
and
denoting the mean and absolute value, respectively, the difference can be quantified in summary by the mean relative deviation of URANS from LES regarding the statistical moments in
Table 3 with
and
, which are
and
respectively. From this, good agreement of both approaches becomes visible. However, when considering the probability density functions, correspondence and deviations between LES and URANS become clear in more detail. A part of the interpretation of the results is based on the fact that, if the first-order Sobol index
is large with respect to an input variable and the response function additionally shows linear behavior regarding this parameter, then the result variable approximately shows a similar distribution type
, as defined for the input variable. Considering the affine model for the approximation of a single response (Equation (
32)), for which the variance is predominantly determined by one input variable, the shape of the probability density function after propagation remains approximately the same whereas the expectation and variance change. From this, conclusions can be made and the plausibility of the results can be checked.
Starting with
and
in
Figure 13a,b, it is noticeable that the shape of the PDF is similar to a normal distribution. This is the case because the variance is predominantly caused by the temperature specification at the left and right walls via
, as can be seen from the first-order Sobol indices
in
Table A3, and shows a nearly linear behavior that becomes visible due to the low interaction behavior through the second-order Sobol indices
in
Table A4 and a low
in
Figure 9e,f. The distribution for
tends towards a log-normal distribution. As can be seen in
Table A3, the variance is predominantly determined by the wall-tangential temperature gradient via
, which was specified with a log-normal distribution. For this reason, a similar distribution for the result originates. For
, the effects of
also become visible. Increasing
enlarges the amount of helium at the top wall. Due to the lower density of helium, this attenuates the erosion at the top wall. This results in a lower convective heat transfer in this region. This creates the superimposed distribution based on the effects of
and
, shown in
Figure 13d. When comparing LES and URANS, the probability density distributions in
Figure 13 and the statistical moments in
Table 3 are in very good agreement. If taking LES as a reliable reference, the absolute value of the Nusselt number for URANS is slightly overestimated on the left, right, and top walls.
The distributions for
and
in
Figure 14a,b, approximately resemble the shape of a chi-distribution because the response function evaluates the absolute difference between the mixing transients. Analogous to
, the variance primarily arises through the variance in the temperature at the left and right walls, which is normally distributed and dominates the distribution of the response. This can be seen in
Table A3 by means of the first-order Sobol indices
. The distributions of URANS and LES are in good agreement. The distributions for the top and bottom walls in
Figure 13c,d and
Figure 14c,d are nearly the same as that for
because the top and bottom walls are adiabatic for the reference case, and this results in the integral deviation from the abscissa.
In
Figure 15a, a normal distribution also arises for the PDF of
, as again the temperature on the left and right walls via
causes the major variance, as can be seen with
in
Table A3. The global kinetic energy
is underestimated with URANS and leads to longer mixing times
in
Figure 16a since the mixing process is dominated by the arising natural convection. The distribution for LES is wider than for URANS because LES captures most of the turbulent structures. The larger amount of the standard deviation also explains the larger deviation from the reference case
for LES in
Figure 15b. By taking the LES as reference, the results for
indicate that URANS is able to predict the variability of
at
due to the parameter uncertainties.
In
Figure 16a, the shape of the PDF for
resembles a normal distribution due to the large influence of
for both LES and URANS, as can be seen with
in
Table A3. Low values for
in
Figure 12d and
in
Table A4 indicate the approximately linear behavior of
. In comparison to URANS, LES predicts shorter mixing times and causes a smaller width of the distribution, which also can be measured by the smaller mean value
and standard deviation
in
Table 3. The distributions for
in
Figure 16b tend towards larger values, as can be seen especially for the LES results. This tendency arises from the truncated normal distribution for the linear stratification parameter
, which increases the segregation intensity and causes the largest proportion of variance, as indicated by
in
Table A3. The variance component including the smaller values mainly arises from the remaining parameters
and
. In contrast to the larger integral mean
of the segregation intensity
in URANS, the results indicate that LES has a higher mixing intensity
whereas the width of the distributions is almost identical. The shorter mixing time measured by
for LES and the clear difference in
between LES and URANS arises since the more resolved convection in LES, which might not be represented sufficiently accurate by the turbulence model in URANS, results in a more accurate prediction of mass transfer, since anisotropic large-scale eddies have a large influence on the mixing process. The deviation from the reference case
, which provides a good measure for the variability of the results when the parameters change, shows very good agreement between LES and URANS.