2.1. Basic System for Compressible Reacting Flow
Our research is based on compressible, spatially filtered Navier–Stokes equations with Boussinesq closure for turbulent scalar transport. The equations read:
In the systems (1)–(4), the summation over repeated indices is adopted. The overbar denotes spatial averaging across the filter of size
, whereas the tilde denotes Favre spatial averaging across the same filter, as it is proposed in [
25]. Unlike Reynolds averaged correlations, for the spatially filtered correlations, the Germano treatment is used:
Analogous treatment is applied to correlations that are spatially filtered according to Favre rules ().
The gas is considered as a mixture of species, so it is necessary to solve transport Equation (4) for mass fractions . In fact, it is possible to solve only equations because the sum of is equal to 1; nevertheless, we prefer not to exclude any mass fraction equations to prevent the accumulation of numerical errors. This approach implies the use of a re-normalization procedure to keep the sum of equal to unity at each time step of the simulation.
Equations (2) and (3) contain the sum of subgrid and molecular stresses, which is represented as
where
is a rate of strain tensor and
is a dynamic molecular viscosity, defined by Sutherland’s formula:
Energy Equation (3) includes total enthalpy, which is denoted by
and represents the sum of specific enthalpy of species formation
, sensible enthalpy
and kinetic energy, so that
. For the summary enthalpies of separate components,
, approximations by quadratic polynomials are used, based on the database [
26] and applicable within the range
. Here
is the subgrid-scale turbulence energy. Total energy
equals
. Inclusion of species formation enthalpy into
and
allows to exclude the chemical source term on the right-hand side of the energy equation. The specific heat of a mixture at constant pressure is calculated as
. Energy and species transport equations utilize turbulent subgrid viscosity
for the closure of scalar-velocity correlations
and
, which involve turbulent Prandtl and Schmidt numbers. Hereinafter, these parameters remain constant and equal to 0.9 and 1.0, respectively, although their influence was considered for LES in [
27] and for round jets in RANS simulations presented in [
28,
29]. The energy equation also contains the work produced by subgrid and molecular stresses.
Velocity correlations
represent the subgrid-scale momentum transport and can be described by either the Smagorinsky model or the more complicated DSM approach, the latter involving six additional differential equations for the components of the subgrid stress tensor
. To simulate compressible flows with combustion, the DSM model [
1], which was originally calibrated against an incompressible database, required several modifications to achieve stability. These modifications include compressibility correction, anisotropy invariant stabilization, subgrid-scale diffusion simplification, and numerical smoothing. All these techniques are described in more detail below. The updated model was tested on decaying turbulence, and its statistical characteristics remained consistent with theoretical requirements.
The chemical source for the
-th mass fraction equation
requires separate consideration due to the influence of turbulence on the mean reaction rates, also known as turbulence-combustion interaction, or TCI. The first channel of TCI, which affects reaction rates due to turbulence fluctuations, results in perturbations of the flame front, affects the formation of premixed areas in recirculation zones, and occasionally can cause fire extinction [
30]. The main contribution into
is made by small-scale turbulence that is not resolved within the LES approach. In the present paper, this contribution is simulated with the use of the steady non-premixed flamelet model [
23] based on the assumption of high Damköhler numbers.
2.2. Chemical Kinetics and Flamelet Library
For methane-air combustion, the considered species are
,
,
,
, and
. The process of combustion is represented by a global kinetic mechanism by Hu, Zhou and Luo [
21,
22] with a single unidirectional reaction:
This simple reaction mechanism does not contain the process of radical formation; it is incapable of reproducing the correct induction time, so it assumes that the flame is already ignited. The rate of the reaction (5) is approximated by the following formula [
21,
22]:
(no summation over
), where
is the molecular weight of the
j-th component and
is the molar stoichiometric coefficient of the
j-th component in the reaction. In Equation (5) (
,
,
,
),
is the universal gas constant,
, and
.
The steady non-premixed flamelet approach is based on the filtered mixture fraction variable
and its variance
. The mixture fraction [
23] is a variable that describes the mixing. It is zero in pure air and one in fuel and varies between these limiting values.
To determine
and
, one should solve two additional differential equations. The first of them is a simple transport equation for a passive scalar, while the second contains two source terms, production
and scalar dissipation
:
Equation (7) requires closures for turbulent transport terms
and
. They are approximated using turbulent subgrid viscosity
to maintain compatibility with the Smagorinsky model (the latter will participate in the comparison of subgrid models):
Here
for simplicity. Closures for source terms are standard and are based on
, filtered gradients for production, and on the subgrid turbulent timescale
(
) for scalar dissipation:
The value of
was found in [
1], and the value of
was determined in the present study as a result of numerical optimization in a series of preliminary simulations with the Smagorinsky subgrid model. The value of the resolved (i.e., spatially filtered) reaction rate
in Equation (4) is calculated using the assumption (which is similar to an ergodic hypothesis) that spatial filtering is equivalent to integration with the probability density function:
where
is the kernel of the spatial filter with characteristic scale
, and
is the probability density function of mixture fraction
. Within this paper, a presumed beta PDF [
1,
23] is used that depends upon two local parameters:
and
. Additionally, in Equation (9),
is the local spatially filtered pressure, and
is the local conditionally filtered value of scalar dissipation:
where
is the Dirac delta function and
is the value of the mixture fraction, for which the stoichiometric ratio of diffusive mass fluxes of fuel and oxidizer is reached [
23]:
Here is the mass stoichiometric coefficient, and and are the mass fractions of fuel (methane) at and oxidizer (oxygen) at , respectively.
Within this paper, the Pitsch and Peters hypothesis [
31,
32] concerning the distribution of the scalar dissipation
is used, which goes back to the Kolmogorov and Oboukhov 1962 theory [
33] on the turbulent kinetic energy dissipation distribution. According to Pitsch and Peters, scalar dissipation obeys similar rules and has a lognormal distribution, which means that the logarithm of
has a Gaussian distribution.
Pitsch and Peters have proposed the following approximation of the lognormal distribution of scalar dissipation:
It is possible to express
from Formula (11). Averaging of this expression and neglecting the contribution of
leads to the following simplified formula for
that was used in our simulations:
The optimal value of the constant was determined in a series of parametric calculations; it differs from the value stated in (8). Note that a significant influence on the levels of temperature was found. This value will be utilized for both DSM and Smagorinsky simulations.
This simplified model for the stoichiometric scalar dissipation rate utilizes cell averaged values, unlike its original definition (10), where the averaging is performed only over the stoichiometric mixture locations inside the filtering region. In the works [
31,
32], it is taken into account by assuming some distribution of
and establishing the relation between
and
. However, the question about the validity of the assumed distribution for the considered flow arises. A simplified model (12) can lead to the amplification of flame-turbulence interactions. The authors plan to search for an adequate model and study its influence on the simulation of the presented flow in the future.
Dependencies
and
in (9) are determined from the numerical solution of the steady one-dimensional flamelet equations [
1,
23]. For this purpose, the in-house code
zFlare-flamelet (TsAGI) was applied. Calculations were performed for 10 values of
and for 10 values of
. After that, for each pair
the integrals (9) were calculated for 100 values of
and 50 values of
. Thus, a four-dimensional flamelet library of size
was obtained.
Due to the fact that this flamelet library describes fire extinction for high
values, the flame front may reveal the presence of extinction holes. Moreover, there could be self-ignition combustion in premixed areas of recirculation. In these cases, the model of steady, non-premixed flamelets can give inadequate reaction rates. To avoid unphysical behavior, an original reaction rate limitation is proposed in this article, which involves bounding by a quasi-laminar reaction rate taken at a temperature
obtained from the flamelet library:
The parameter is calculated using PDF, analogous to (9), whereas is the local value of the resolved temperature obtained from the solution of Equations (1)–(4). The chemical source (9), averaged using the flamelet library, is usually smaller than the right-hand side of (13), which is based on local parameters. Therefore, the restriction (13) works only in rare situations where the model of steady, non-premixed flamelets is inapplicable. For example, sometimes it is observed that the fuel (or oxidizer) is totally consumed on the stoichiometric iso-surface, while the flamelet-based formula (9) predicts a finite rate of the reaction. Such situations can arise because flamelets are used only for the calculation of chemical sources, whereas concentration fields are obtained from Equation (4) and are not strictly coupled with the mixture fraction . The flamelet library provides source terms based only on pressure, scalar dissipation rate , mixture fraction and its variance . It is not capable of detecting such situations, whereas the restriction (13) solves this problem.
It is necessary to consider the possibility of using the kinetic scheme from [
21] with the TCI approach described above. As stated in the article [
34], this kinetics by Zhou is based on Westbrook’s 1981 set of 3 kinetics [
35] for premixed methane-air flames. It uses the same value of
, the same powers of concentration, although kinetics by Zhou has a higher pre-exponential multiplier and neglects density.
Testing of this kinetics in laminar flame calculations revealed that it significantly overestimates laminar flame speed, predicting values of
near stoichiometric mixture depending on the equivalence ratio, whereas the experimental values are close to
. However, it is well known that flamelet models of TCI have a low sensitivity to the chemical kinetics. To check the possibility of using the kinetics from [
21], the flamelet library data obtained with this kinetics was compared with analogous data obtained using the four-reaction quasi-global kinetic scheme of methane-air oxidation by Basevich et al. [
36]. The latter was verified to give an appropriate laminar flame front velocity for the equivalence ratio below 1. The third kinetics in this comparison is BFER2 kinetics by Franzelli et al. [
37], which was developed to describe partially premixed flames and to represent methane-air combustion in a wide range of equivalence ratio values, which is justified according to its representation of laminar flame speed. Because different reactions are used in these kinetic schemes, we compared the cell-averaged source for
. The results are given in
Figure 1.
In the simulations presented below, the time-averaged value of scalar dissipation at the stoichiometric surface is
. According to
Figure 1, for these values of
the flamelet model predicts close values of the average chemical source with all considered kinetic schemes.
The largest differences in prediction of the cell-averaged source with these kinetic schemes can be found at high values of , and they affect the description of the extinction. In the case of Zhou kinetics, extinction occurs at a very high value , which is almost not found in calculations. The Basevich et al. kinetics show complete extinction at . The Franzelli et al. kinetics has the lowest extinction value of the scalar dissipation rate, which is equal to . To understand the role of extinction, the comparison between Zhou kinetics and BFER2 kinetics in the simulations of the swirling burner will be provided below.
2.3. DSM Model Formulation and Modifications
In the DSM approach, six differential equations for the subgrid stress tensor components are solved:
It requires closures for turbulent transport
, for the pressure-strain tensor
, which is responsible for the turbulence kinetic energy redistribution between the Cartesian components, and for the dissipation rate tensor
. The latter tensor can be considered isotropic,
, in developed turbulent flows [
38]. Here
is the dissipation rate, which is calculated in the current article according to [
1].
Compared to [
1], several modifications were introduced in the model to account for the compressibility effects in the considered flows.
The first improvement affected the triple velocity correlation term. It turned out that the previously proposed model may be insufficiently stable in compressible flows, so we added a term representing the simple gradient diffusion with the weight
to the
model:
This modification completely removes “implosion instability”. Under “implosion instability”, we mean amplification of the disturbances in the rarefaction zone near the methane injector observed in the preliminary simulations without the modification. This effect may occur due to weak rarefaction, when anisotropic turbulent transport is unable to suppress the growing velocity gradients, so subgrid-scale parameters experience unphysical behavior by amplifying rarefaction during the implosion. The key difference between the additional term and the previously introduced -term is that turbulent subgrid viscosity depends on the strain tensor norm , consequently it is connected to the resolved field, whereas the -term adopts “turbulent viscosity” based on subgrid parameters only.
The second improvement is related to accounting for compressibility in the pressure-strain correlation closure. The basic representation of the closure is as follows [
1]:
with tensor multiplication obeying the matrix rule,
. Here
is a unity tensor,
and
are the resolved (deviatoric) strain rate and rotation tensors, respectively.
Hereinafter,
,
and
denote the anisotropy tensor and its invariants:
In compressible flows, the strain tensor closure should also contain the separate terms “linear” in the divergence of velocity and “anisotropy tensor”
. The form of this term was considered, e.g., in [
39], where it was calibrated to obtain the correct turbulent kinetic energy downstream the shock wave:
This term was incorporated into the present model, but due to the fact that we are simulating a subsonic flow, the coefficient is different and equals .
The third modification implies anisotropy invariant stabilization, which bounds the Lumley’s invariant by a near-zero value to prevent possible numerical instabilities near a two-component limit of turbulence. Moreover, this improvement also helps to have an adequate probability density function for the invariant, as is illustrated below.
To understand the working principle of this fix, let us consider a case of small
falling within the range
. Then the subgrid-scale stress tensor is shifted towards the isotropic form,
The value of the “isotropization factor”
is obtained according to the formula
where
. This choice of
guarantees that after the isotropization,
.
The value of
is chosen to be 0.1. To demonstrate the effect of this improvement, we show the one-point probability density function of
in a series of isotropic decaying turbulence simulations (see
Figure 2). The blue line represents the PDF before the limitation, and the black line represents the PDF after the limitation. In this figure, the peak near the maximum allowed value of
vanishes after implementing the anisotropy invariant stabilization.
The fourth modification to this model is necessary for simulations on non-uniform grids. It introduces a modification to the numerical scheme to make it more dissipative. The problem is that the DSM approach utilizes central differences for convective term approximation. Depending on the flow regime, this numerical scheme may produce even-odd decoupling. One of the ways to avoid this problem is to use asymmetric one-sided gradients for the source terms of turbulent parameters. Although it removes oscillations with high wavenumbers for a uniform grid, it may have problems with non-uniform grids. Here arises one more problem, which is connected to a commutation error when the filter size gradient exists. These two factors make the representation of the highest resolved frequencies highly inaccurate, especially in the cells stretched in one direction. In this case, it would be better to have a more dissipative numerical scheme to suppress the oscillations, which are usually four or three cells in wavelength.
The simple solution is numerical smoothing, which subtracts the error of two different approximations of the strain rate tensor in the momentum equation. In other words, it is an additional numerical flux
which is added to the momentum equation as follows:
Here, the square brackets denote the gradients calculated at the cell sides by using values of from the centers of the adjacent cells, whereas the angle brackets represent the gradients, which were calculated by interpolating the values from the centers of the adjacent cells. In our simulations, .
On a uniform grid, this modification introduces numerical smoothing of third order with respect to cell size, . It also complies with the requirement of preserving the scheme’s conservativity.
To check all these modifications and make sure the final formulation of the DSM model represents the basic turbulence statistics properly, test simulations of decaying isotropic turbulence were performed. A grid of cells in a periodic box of size was used with an integral turbulent length scale of . To specify the initial velocity field, the box was filled with harmonic waves according to the von Karman spectrum. The simulations were performed until the resolved energy reached one tenth of its initial value, which was taken as .
In
Figure 3, the comparison between the new, more dissipative method and the basic method shows that the longitudinal energy spectrum
starts to deviate from Kolmogorov’s “–5/3” law at wavenumber 30, whereas the basic method starts deviating at wavenumber 40. This tendency emphasizes that the waves with wavelengths equal to four cells are still present in the simulations with the basic version of the model, whereas with the modified version, these parts of the spectrum have already started to dissipate. In addition, the inertial subrange of the spectrum remains close enough to Kolmogorov’s law. This allows us to conclude that all the modifications do not impair the description of isotropic turbulence.
2.4. Formulation of the Test Case
The test case represents a non-premixed, unconfined flame in a swirl burner. The data were obtained by two research groups: the velocity measurements were performed by Yasir Al-Abdeli at the University of Sydney, whereas the temperature and species concentrations were measured at Sandia National Laboratories by Dr. Rob Barlow and Dr. Peter Kalt. These two groups had the same burners, although the wind tunnels providing the co-flow were different. Sandia National Laboratories had a cross-section measuring , whereas the research at Sydney University utilized a smaller wind tunnel with a cross-section of . Peter Kalt and Rob Barlow investigated concentrations and temperature, so they used pure as fuel, whereas Yasir Al-Abdeli considered the combustion of CNG, which contained of and of butane and other hydrocarbons, although this should not significantly affect velocity measurements. The burner consists of a bluff-body surrounded by an annular shroud measuring in diameter. The bluff-body face measures in diameter and has a central fuel inlet, that is in diameter. The annular shroud has tangential inlets to control swirling. There are three inlets downstream: the honeycomb, which is made of wire mesh, and the flow straightener. The latter is applied due to the perpendicularity of the main air inlets and the annular shroud.
The experimental database contains three types of test cases. The cases marked with SM are non-premixed methane flames; the cases labeled SMH contain hydrogen in combination with methane as fuel, so these cases require more complicated kinetics. Finally, the cases designated SMA represent partially non-premixed combustion, as stated by Yasir Al-Abdeli, so the simulations of these cases using the mixture fraction-based flamelets may be inadequate.
Our simulations were focused on the case SM1, as we decided to compare our results to the data obtained by Hu, Zhou, and Luo [
21], who applied the Pressure Implicit Splitting Operator algorithm to perform the pressure-velocity correction in their solver.
The case SM1 has a fuel axial velocity of , axial velocity of external swirled air of , tangential velocity of . This results in a swirl number equal to 0.5. The co-flow longitudinal velocity is . The visual flame length is estimated at approximately . The Reynolds number for the swirled external jet is , whereas for the internal fuel jet, it is .
To perform an LES, it is necessary to use computational grids with cell sizes that correspond to the inertial interval of the turbulence. Unless resolving the boundary layer on the bluff-body face is required, it is enough to take at least one tenth of the integral turbulence length scale as the size of a cell. Taking into account our computational capabilities, we chose approximately 200 cells in the longitudinal (-axis) direction with sizes close to .
Our computational grid, which is shown in
Figure 4, is
cells in the
,
, and
directions, respectively. To describe the central jet, 20 cells were placed across its diameter, so the grid has refinement near the central axis. The computational domain itself is
. It has a square shape to avoid grid skewness. The round shapes of inflowing jets were approximated with a second order of accuracy using the injector boundary condition described in [
40].
The boundary conditions are the following: subsonic inflow for the central fuel inlet; annular shroud inlet and co-flow; constant pressure outflow at the outlet ( was specified to prevent inflow); and symmetry planes for the bluff-body face and side surfaces. “Subsonic inflow” means solving the Riemann problem at boundaries. To avoid reflection of acoustic waves from side planes, it is better to make the external part of the grid coarse to suppress them with higher dissipation. After trying the Riemann-type boundary condition on side surfaces, it was found to be unreliable due to the proximity to the jet because it can reverse the flow and lead to a qualitatively different solution.
The grid contained 3.145 million cells. The time step for the PDF flamelet model was limited to
, whereas in our program, quasi-laminar simulations typically require the time steps of
to replicate the chemical reactions correctly. The time required for the central jet to traverse the flame is approximately
, so the simulation time interval should be about
. It was found that the bottleneck-shaped structure of the flame becomes stable after
, although after
it slowly disappears, presumably due to the interaction with the side boundary conditions. This process may occur due to the lack of ejection into the swirling region, which is partially suppressed in our simulation by the impermeability of the side walls. However, there exists an almost statistically steady state before the collapse of the bottleneck structure. We chose our primary time averaging interval in the middle of this state, between
and
. We also continued our simulation up to
. Further simulations revealed that the flow evolves into a completely different conic-type structure of the flame. A comparison of the time-averaged pressure fields before and after the breakup of the bottleneck is shown in
Figure 5. The rarefaction zone appears near the side boundaries, and while it flows downstream, it slowly destroys high pressure zones near the bottleneck. This process takes place in both Smagorinsky and DSM simulations, so its dependence on the subgrid-scale stress model is weak. Thus, we conclude that it takes place mainly due to interactions of the acoustics with the boundaries.
Hereinafter, all the provided results will refer to the flow structure observed before the bottleneck breakup, namely between 13 and 19 ms. Time averaging was applied from 14 to 17 ms. During the averaging period, azimuthally-averaged instant temperature fields were found to be stable.