Researchers working with the
OpenFOAM
software
package [1] are concerned with comparing accuracy and efficiency across various
solvers. Using
OpenFOAM
for solving computational
fluid dynamics (CFD) problems involves selecting an appropriate solver that
meets the expectations of accuracy, efficiency, and stability. To achieve this,
it is necessary to
analyse
the performance of the
solver based on some reference solutions. However, few studies are offering a
detailed comparison of
OpenFOAM
solvers based on
these characteristics. Some examples of such comparisons can be found in [2]
and [3], but they do not provide clear recommendations for solver selection. A
more detailed attempt at comparing solvers was made in a series of works where
the reference problems included supersonic flow around a cone at an angle of
attack [4, 5], shock wave modelling [6], and the formation of a two-dimensional
rarefaction wave [7]. A review of these studies can also be found in [8]. The
method proposed in these works gained popularity in [9], [10]. By employing the
generalised
computational experiment technique
[11-14], results were obtained that allow users to navigate the wide range of
numerical methods developed and select the most accurate and efficient ones for
their calculations. The
generalised
computational
experiment involves dividing the governing parameters of a problem within a
specific range, followed by a parametric study and
visualisation
of multidimensional results [15].
OpenFOAM
(Open Source
Field Operation
And
Manipulation CFD Toolbox) is a
popular software package for computational fluid dynamics (CFD), widely used
for modelling liquid and gas flows, heat transfer, and a variety of other tasks
related to physical processes.
OpenFOAM
is an
open-source package, meaning that it is available for free, and its source code
is open. This allows users to modify the code for their specific tasks and add
custom equations, turbulence models, or other physical processes. For example,
if you have a specific task, such as modelling gas-dynamic shock waves with
chemical reactions, you can independently create a solver that takes these
features into account. This flexibility distinguishes
OpenFOAM
from commercial software, where many features are closed to modification.
OpenFOAM
also supports parallel computing, enabling the
solution of large-scale problems with complex physics. This is critical when
calculations need to be performed on a grid consisting of millions of cells. In
OpenFOAM,
the computational domain can be easily
divided among multiple processors, thereby accelerating the calculations. The
comparison included four solvers: two standard solvers –
rhoCentralFoam
and
sonicFoam
and two custom solvers –
pisoCentralFoam
[16] and
QGDFoam
[17]. The latter two solvers were developed by research teams from the
Institute for System Programming of the Russian Academy of Sciences and the
Keldysh
Institute of Applied Mathematics of the Russian Academy
of Sciences.
The
rhoCentralFoam
solver
utilises
a central-upwind scheme of the Godunov type, originally proposed by
Kurganov
and
Tadmor
(KT) [18],
later modified in the work by
Kurganov,
Noelle, and
Petrova
(KNP) [19]. The
Kurganov-Tadmor
scheme bypasses the Riemann problem by using an approximation of local wave
propagation speeds and a central approach. Instead of solving the Riemann
problem to determine the flux at the cell interface, the KT scheme employs an
estimate of the maximum wave propagation speed (typically determined through
the maximum characteristic speed of the equations) on both sides of the cell.
This speed corresponds to the rate at which information propagates through the
cell boundary. For example, in the case of the Euler equations, this might be
the speed of sound plus the flow velocity, as this is the maximum speed of
information propagation in a compressible gas. This approach helps to avoid
many difficulties associated with computing shock waves and steep gradients in
density or pressure fields, which can lead to oscillations in numerical
solutions. The KNP scheme was implemented for
OpenFOAM
by
Greenshields
[20].
The
sonicFoam
solver uses the
PIMPLE algorithm [21] to link pressure and velocity. This algorithm
combines elements of the PISO and SIMPLE methods. At its core, the PIMPLE
algorithm features an iterative structure designed to refine the velocity and
pressure fields within the fluid domain. The algorithm begins with an initial
assumption about the velocity field, which is often derived from previous time
steps or estimated using simpler approaches. This initial assumption is crucial
as it lays the foundation for the subsequent iterative process. In the first
step, the algorithm predicts the velocity field based on the existing pressure
distribution, applying an explicit scheme that facilitates direct calculation
of the velocity. After obtaining the initial velocity forecast, the algorithm
then updates the pressure field. This is where the strength of the PISO method
comes into play. At this stage, the algorithm calculates a pressure correction
that ensures compliance with the conservation of mass. The pressure correction
is essential as it helps eliminate discrepancies between the predicted and
actual
behaviour
of the gas, especially in regions
where flow velocities can change rapidly or where complex boundary conditions
exist. After correcting the pressure, the velocity field is updated once again,
taking into account the new pressure information. This iterative process is
repeated several times, typically within a single time step, allowing for
continuous refinement of the velocity and pressure fields. Each iteration helps
improve the accuracy of the solution, leading to greater stability.
When using explicit methods based on solving the Riemann problem or flux
vector splitting, the following issues arise: at low Mach numbers, the solution
either loses stability, or an extremely small time step is required, which
increases computational costs and makes such methods inefficient for the
considered problems. On the other hand, implicit splitting methods such as PISO
and SIMPLE can cause non-physical oscillations in the numerical solution and
loss of stability at high Mach numbers. These phenomena significantly limit the
effectiveness of these methods in problems where compressibility plays a major
role, making them unsuitable for studying high-speed flows or acoustic
phenomena. The solution could be to use a hybrid approach, where PISO/SIMPLE
algorithms for the implicit integration of mass, momentum, and energy
conservation equations are combined with non-oscillatory methods for
discretizing convective terms. The authors of [16] selected the
Kurganov-Tadmor
scheme as the non-oscillatory method; it is
already implemented as an explicit scheme in
OpenFOAM,
has undergone successful testing, and is simple enough for integration into a
hybrid approach. One of the key advantages of this scheme is the independence
of flux approximation for physical quantities from the system of equations'
characteristics, eliminating the need to find Riemann invariants and decompose
solutions by characteristics. The main idea of the hybrid method is to
introduce a switching function that translates the flux approximation from the
KT scheme to the form used in the PIMPLE method, depending on the proximity to
subsonic speeds [22]. All these hybrid solvers are not included in
OpenFOAM, but are available in the authors' public
repository [23].
The quasi-gas-dynamic (QGD) equations system was developed by a team led
by B.N.
Chetverushkin
[24]. It is based on the idea
that gas can be considered not as a continuous medium, but as a set of
particles moving and interacting with each other at the microscopic level. The
QGD model serves as an intermediate approach between classical hydrodynamics
and more complex models that account for microscopic effects. The introduction
of additional terms into the quasi-gas-dynamic equations aims to capture
small-scale effects that cannot be directly resolved within the framework of
traditional Navier-Stokes equations [25]. These
additional terms are often referred to as dissipative terms, and they are
introduced through various averaging and approximation techniques that reflect
physical processes. The inclusion of these terms allows for a more accurate
description of complex phenomena and interactions in gases, especially in
conditions where non-equilibrium processes and energy transformation are
significant. Based on this system, the
QGDFoam
solver
was developed. The parameter
τ
plays
a
key role in accounting for small-scale effects and determines the time scale
for dissipative processes. The controlled parameter
α
(associated
with
τ),
in dissipative terms, provides this solver with adjustable numerical viscosity,
which helps to reduce unwanted oscillations at discontinuities.
In this study, a two-dimensional inviscid problem of steady flow
formation around a blunt circular cone with a half-angle of
β
is
used to compare solvers. The flow is generated by a
supersonic gas flow with a Mach number
М
at zero angle of attack. The
defining parameters of the problem, in terms of the
generalised
computational experiment, are the Mach number
М
and the half-angle of the cone
β.
The ranges for the varying parameters and their increments were chosen the
following way: the Mach number
М
is set to 2, 4, and 6, and the
half-angle of the cone
β
is
set to 5°, 10°, 15°, and 25°. The overall flow scheme is presented in Figure 1.
For all selected parameters, there exists a tabulated solution for the problem
[26]. The Euler equations system, closed by the ideal gas equation of state, is
taken for the calculations.
Figure
1.
Flow
Diagram
The computational domain, as shown in Figure 2, is divided into cells.
To work with the
OpenFOAM
package, it is necessary to
define the boundary and initial conditions. At the inlet boundary, designated
as "inlet," the characteristics of the undisturbed flow are
specified: the pressure P is 101,325 Pa, the temperature
T
is
300 K, and the x-component of velocity
Ux
varies from 694.5 m/s (2 Mach) to 2083.5 m/s (6 Mach), while the y-component of
velocity
Uy
is 0 m/s. At the outlet and the
top, conditions are set so that the derivatives of gas dynamic functions normal
to the boundary are equal to zero, defined in
OpenFOAM
as «zeroGradient». At the cone boundary, a zero
gradient condition is applied for pressure and temperature, while a «slip»
condition is used for velocity, which corresponds to the non-penetrating
condition in the Euler equations. To model the axisymmetric geometry in
OpenFOAM, this special «wedge» condition is applied at the
front and back boundaries. The axis in
OpenFOAM
uses
a specific boundary condition «empty», which is employed when calculations in
that direction are not conducted.
The computational domain (Figure 2) is divided into cells. The
OpenFOAM
package requires the definition of the boundary
and initial conditions for the solution. At the inlet boundary, the parameters
of the undisturbed incoming flow are specified (pressure P = 101325 Pa,
temperature T = 300 K, the x-component of velocity varies from
694.5 m/s (2 Mach) to 2083.5 m/s (6 Mach), and the y-component of velocity is 0
m/s). At the outlet and the top, boundary conditions are set such that the
derivatives of gas dynamic functions normal to the boundary are equal to zero,
defined in
OpenFOAM
as «zeroGradient».
At the cone, a zero-gradient condition is also applied for pressure and
temperature, while a «slip» condition is used for velocity, corresponding to
the no-penetration condition in the Euler equations. For modelling the
axisymmetric geometry, a special «wedge» condition is applied at the front and
back boundaries. This special boundary condition called «empty» is applied at
the axis. This condition is used in cases where calculations are not performed
in that direction.
The scheme of the computational domain for the cone with a half-angle
β
= 15° is
presented in Figure 2. It is worth noting that in the displayed image, the grid
appears larger than in the actual calculations for clarity.
Figure 2. Computational Domain Scheme
The number of grid cells is 35,950. The initial conditions correspond to
the boundary conditions at the inlet face, meaning the parameters of the
incoming flow serve as the initial conditions. In the
QGDFoam
solver, a smoothing coefficient of
α
= 0.1 was applied across
the entire computational domain. Additionally, the values for molar mass M =
28.96 and specific heat capacity at constant pressure
Cp
= 1004
were set.
Unlike many other software packages,
OpenFOAM
manages simulations through text files. This provides flexibility, as it is
easy to automate task launches, modify simulation parameters, and process
results. The
standardisation
of calculations plays a
crucial role in
organising
the comparison of solvers,
as it ensures uniform conditions for assessing their performance and accuracy.
With
standardised
methodologies, grids, boundary
conditions, and physical models, the results become comparable and reliable.
This allows researchers to eliminate the influence of external factors and
focus on the characteristics of each specific solver. Furthermore,
standardisation
helps in better understanding the strengths
and weaknesses of each solver, which can be useful for selecting the optimal
tool based on the specifics of the task. In the
OpenFOAM
package, we used the same parameters for the configuration files
fvSchemes
and
fvSolution
as in
[5].
The simulations performed with all solvers resulted in the well-known
qualitative flow pattern for the considered problem. An example is shown in
Figure 3, depicting the pressure distribution in the computational domain. The
presented pressure distribution was obtained using the
rhoCentralFoam
solver. No solution breakdown was observed for any of the solvers, indicating
the high
stabilising
properties of all solvers
involved in the study.
Next, the deviation is evaluated from the exact solution for the entire
computational domain using the L2
norm. To do this, we define the
relative error Err for the L2
norm as follows:
here,
y
mexact
represents the pressure
p,
density ρ,
x
and
y
components of velocity
(Ux
and
Uy).
Vmexact
is the cell volume. The values of
ymexact
were
obtained by interpolating the
tabulated solution of the problem. The solvers involved in the comparative
accuracy analysis were
sonicFoam,
QGDFoam,
rhoCentralFoam,
and
pisoCentralFoam.
For convenience, the solvers are denoted by abbreviations in the tables:
rCF
(rhoCentralFoam),
pCF
(pisoCentralFoam),
sF
(sonicFoam),
QGDF
(QGDFoam).
The deviation values from the exact
solution for all values
over the entire computational domain are
provided in Tables 1–3. The smallest values in each row are highlighted in
bold.
Figure
3.
Field
steady-state flow pressure for the
rhoCentralFoam
Solver
TABLE
1
ERRORS
FOR
M = 2
Value
|
Half-cone
angle
|
rCF
|
pCF
|
sF
|
QGD
F
|
p
|
5
|
0.017028
|
0.022472
|
0.031921
|
0.025685
|
10
|
0.027280
|
0.033699
|
0.055463
|
0.042351
|
15
|
0.034960
|
0.040031
|
0.080514
|
0.046906
|
25
|
0.036960
|
0.043399
|
0.085460
|
0.049201
|
ρ
|
5
|
0.013049
|
0.016927
|
0.026125
|
0.019875
|
10
|
0.021151
|
0.025949
|
0.046866
|
0.033427
|
15
|
0.027018
|
0.030899
|
0.069235
|
0.037590
|
25
|
0.028358
|
0.033815
|
0.074514
|
0.040236
|
Ux
|
5
|
0.003332
|
0.002882
|
0.003195
|
0.003090
|
10
|
0.005847
|
0.005424
|
0.006937
|
0.005807
|
15
|
0.009694
|
0.008440
|
0.012859
|
0.009248
|
25
|
0.012425
|
0.012269
|
0.016679
|
0.012820
|
Uy
|
5
|
0.049208
|
0.058461
|
0.073065
|
0.061710
|
10
|
0.046843
|
0.054963
|
0.073083
|
0.060938
|
15
|
0.043771
|
0.048582
|
0.071450
|
0.048474
|
25
|
0.036296
|
0.044299
|
0.057872
|
0.040218
|
TABLE
2
ERRORS FOR
M = 4
Value
|
Half-cone
angle
|
rCF
|
pCF
|
sF
|
QGD
F
|
p
|
5
|
0.031943
|
0.040018
|
0.067521
|
0.051875
|
10
|
0.045202
|
0.053764
|
0.125038
|
0.062389
|
15
|
0.051245
|
0.060396
|
0.154412
|
0.074341
|
25
|
0.051458
|
0.061273
|
0.161013
|
0.077917
|
ρ
|
5
|
0.024840
|
0.030584
|
0.056221
|
0.040623
|
10
|
0.035139
|
0.041356
|
0.106082
|
0.049628
|
15
|
0.039569
|
0.046781
|
0.134346
|
0.059978
|
25
|
0.039651
|
0.047775
|
0.142780
|
0.063810
|
Ux
|
5
|
0.005091
|
0.003401
|
0.004375
|
0.003689
|
10
|
0.008348
|
0.006382
|
0.011029
|
0.006035
|
15
|
0.011934
|
0.009961
|
0.019059
|
0.009639
|
25
|
0.015016
|
0.014323
|
0.024822
|
0.013115
|
Uy
|
5
|
0.070407
|
0.076146
|
0.105546
|
0.087226
|
10
|
0.060978
|
0.069945
|
0.115991
|
0.065847
|
15
|
0.053252
|
0.063890
|
0.102818
|
0.059047
|
25
|
0.043076
|
0.058077
|
0.081359
|
0.050234
|
TABLE
3
ERRORS FOR
M =
6
Value
|
Half-cone
angle
|
rCF
|
pCF
|
sF
|
QGD
F
|
p
|
5
|
0.057095
|
0.064405
|
0.174251
|
0.083892
|
10
|
0.065626
|
0.067459
|
0.183761
|
0.095616
|
15
|
0.068654
|
0.071534
|
0.194212
|
0.099098
|
25
|
0.071054
|
0.072618
|
0.204263
|
0.100022
|
ρ
|
5
|
0.039118
|
0.044436
|
0.141836
|
0.062416
|
10
|
0.050310
|
0.052647
|
0.158682
|
0.076522
|
15
|
0.053411
|
0.056018
|
0.168282
|
0.080265
|
25
|
0.055427
|
0.057246
|
0.182445
|
0.082623
|
Ux
|
5
|
0.006597
|
0.003779
|
0.007974
|
0.004184
|
10
|
0.010538
|
0.006571
|
0.012771
|
0.007293
|
15
|
0.014176
|
0.009545
|
0.017492
|
0.010283
|
25
|
0.018507
|
0.012650
|
0.024849
|
0.013967
|
Uy
|
5
|
0.090816
|
0.092901
|
0.158190
|
0.101646
|
10
|
0.075070
|
0.077949
|
0.124047
|
0.082505
|
15
|
0.064468
|
0.070916
|
0.097257
|
0.065471
|
25
|
0.056260
|
0.068021
|
0.085289
|
0.055883
|
The error surface visualization shown
in
Figures
4–7.
Figure
4. Variation of the
deviation from the
exact
solution for
pressure
depending on the Mach
number
and
cone
half-angle for
all
solvers in
L2
norm
Figure 5. Variation of the deviation from the exact solution for density depending on the Mach number and cone half-angle for all solvers in L2 norm
Figure 6. Variation of the deviation from the exact solution for velocity x-component depending on the Mach number and cone half-angle for all solvers in L2 norm
Figure 7. Variation of the deviation from the exact solution for velocity y-component depending on the Mach number and cone half-angle for all solvers in L2 norm
The analysis of the presented tables allows us to draw a number of
important conclusions. In general, the errors increase with increasing Mach
number from 2 to 6 for all considered quantities. For pressure and density, the
rCF
method shows the smallest errors in all cases,
and the errors increase with increasing half-cone angle up to a certain value,
after which the growth slows down or stops. For velocity in the x direction,
the
pCF
method shows the smallest errors for all Mach
numbers and half-cone angles, with the errors increasing with increasing
half-cone angle. For velocity in the y direction, the
rCF
method shows the smallest errors in all cases, but unlike the other quantities,
the errors for
Uy
generally decrease with increasing
half-cone angle. The
sF
method consistently shows the
largest errors for all quantities and conditions, while the QGDF method
generally shows intermediate error values between the best (rCF
or
pCF)
and worst
(sF)
methods. It is worth noting that the difference in accuracy between methods
becomes more pronounced with increasing Mach number, especially for pressure
and density. These observations indicate that the choice of the most
appropriate method may depend on the specific problem and the physical quantity
considered, with the
rCF
and
pCF
methods generally performing the best.
However, it should be noted that in this study, the smoothing parameter
α
for the
QGDFoam
solver was not adjusted, and as shown in [27],
modifying this parameter affects the solution’s accuracy.
α
= 0.1
may not be an optimal value for this type of problem. This issue will be
addressed in our future research.
The largest absolute increase in error is observed for pressure at an
angle of 5° for the
sF
method when going from M=4 to
M=6. The error increases from 0.067521 to 0.174251, giving an absolute increase
of 0.10673. The largest relative increase in error at this transition is
observed for the velocity
Ux
at an angle of 5° for
the
sF
method. The error increases from 0.004375 to
0.007974, which corresponds to a relative increase of 1.82 times. It is
interesting to note that for the
Uy
velocity, there
is a decrease in error when going from M=4 to M=6 for some methods and angles.
For example, for the
sF
method at an angle of 15°,
the error decreases from 0.102818 to 0.097257, this corresponds to a decrease
of 5.4%. The
rCF
method shows the most stable results
for both transitions. For example, for pressure at an angle of 25°, the
increase in error is only 39% for the transition from M=2 to M=4 and 38% for
the transition from M=4 to M=6, which is much smaller than the other methods.
This comprehensive study of
OpenFOAM
solvers for supersonic flow around a Spherically Blunted Cone provides valuable
insights into their accuracy across various conditions. A notable trend is the
increase in errors with rising Mach numbers from 2 to 6 for all considered
quantities. For pressure and density, errors typically increase with increasing
half-cone angle up to a certain value, after which the growth slows or stops.
Interestingly, for the y-component of velocity, errors generally decrease with
increasing cone half-angle. The study highlights that the choice of the most appropriate
solver may depend on the specific problem and the physical quantity of
interest. The difference in accuracy between methods becomes more pronounced at
higher Mach numbers, especially for pressure and density calculations. Future
research opportunities include investigating the accuracy of the solvers in
more complex geometries and flow conditions, such as unsteady flows or flows
with strong shock-boundary layer interaction.
This study provides crucial guidance for researchers
and engineers in selecting the most accurate
OpenFOAM
solver for specific supersonic flow problems, emphasizing the importance of
considering both the flow conditions and the physical quantities of primary
interest in their simulations.
Calculations were performed on the hybrid supercomputer K-100 installed
in the Supercomputer Centre of Collective Usage of KIAM RAS.
1. OpenFOAM Foundation: [Online]. URL: http://www.openfoam.org (Accessed: 02.07.2024).
2. Gutierrez L. F., Tamagno J. P., Elaskar S. A. High speed flow simulation using OpenFOAM // Mecanica Computacional. 2012. Vol. XXXI. P. 2939–2959.
3. Lorenzon D., Elaskar S. A. Simulacion de flujos supersonicos bidimensionales y axialmente simetricos con OpenFOAM // Revista de la Facultad de Ciencias Exactas, Fisicas y Naturales. 2015. Vol. 2. no. 2. P. 65–76.
4. Bondarev A. E., Kuvshinnikov A. E. Analysis of the behavior of OpenFOAM solvers for 3D problem of supersonic flow around a cone at an angle of attack // CEUR Workshop Proceedings, 2020, V. 2763, p.48–51, CPT2020, Proceedings of the 8th International Scientific Conference on Computing in Physics and Technology, Moscow region, Russia, November 09-13, 2020.
5. Alekseev A. K., Bondarev A. E., Kuvshinnikov A. E. On uncertainty quantification via the ensemble of independent numerical solutions // Journal of Computational Science. 2020. Vol. 42 101114.
6. Alekseev A. K., Bondarev A. E., Kuvshinnikov A. E. Comparative analysis of the accuracy of openfoam solvers for the oblique shock wave problem // Mathematica Montisnigri, 2019, Vol. XLV. P. 95-105.
7. Bondarev A. E., Kuvshinnikov A. E. Analysis and Visualization of the Computational Experiments Results on the Comparative Assessment of OpenFOAM Solvers Accuracy for a Rarefaction Wave Problem // Scientific Visualization. 2021. Vol. 13. № 3. P. 34–46.
8. Бондарев А.Е., Кувшинников А.Е. Задачи сравнительной оценки численных методов на референтных решениях // Труды Семнадцатой Международной научно-технической конференции «Оптические методы исследования потоков». М.: Научно-технологический центр уникального приборостроения РАН, 2023. C. 517–528.
9. Canteros M.A., Polansky J. Review and comparison of two OpenFOAM® solvers: rhoCentralFoam and sonicFoam // EPJ Web of Conf. 2024. Vol. 299 01005.
10. Analysis of the oscillations induced by a supersonic jet applied to produce nanofibers / Quintero F., Doval A.F., Goitia A. et al. // International Journal of Mechanical Sciences. 2022. Vol. 238. 107826.
11. Бондарев А. Е., Галактионов В. А. Построение и обработка результатов вычислительного эксперимента на основе параллельных решений для оптимизационных и параметрических исследований в газовой динамике (пленарный доклад) // Оптические методы исследования потоков: Труды XIV Международной научно-технической конференции. М.: Издательство «Перо», 2017. С.7–15.
12. Bondarev A. E. On visualization problems in a generalized computational experiment //Scientific Visualization. 2019. Vol. 11.2 P. 156–162.
13. Alekseev A. K., Bondarev A. E., Galaktionov V. A., Kuvshinnikov A.E. On the construction of a generalized computational experiment in verification problems // Matematica Montisnigri. 2020. Vol. XLVIII. P. 19–31.
14. Alekseev A. K., Bondarev A. E., Galaktionov V. A., Kuvshinnikov A. E. Generalized Computational Experiment and Verification Problems // Program. Comput. Soft. 2021. Vol. 47 P. 177–184.
15. Zakharova A. A, Korostelyov D. A., Podvesovskii A .G., Bondarev A. E., Galaktionov V. A. Generalized Computational Experiment State Analysis Using Three-Dimensional Visual Maps // Scientific Visualization. 2022 Vol. 14.4 P. 12–23.
16. Kraposhin M., A. Bovtrikova A., Strijhak S. Adaptation of Kurganov-Tadmor numerical scheme for applying in combination with the PISO method in numerical simulation of flows in a wide range of Mach numbers // Procedia Computer Science. 2015. Vol. 66. P. 43–52.
17. Istomina M.A. About realization of one-dimensional quasi-gas dynamic algorithm in the open program OpenFOAM complex // Preprinty IPM im. M.V.Keldysha. 2018. № 001. [In Russian]
18. Kurganov A., Tadmor E. New high-resolution central schemes for nonlinear conservation laws and convection-diffusion equations // J. Comput. Phys. 2000. Vol. 160. № 1. P. 241–282.
19. Kurganov A., Noelle S., Petrova G. Semidiscrete central-upwind schemes for hyperbolic conservation laws and Hamilton–Jacobi equations // SIAM J Sci Comput. 2001. Vol. 23 P. 707–740.
20. Greenshields C.J., Wellerr H.G., Gasparini L., Reese J.M. Implementation of semi-discrete, non-staggered central schemes in a colocated, polyhedral, finite volume framework, for high-speed viscous flows // Int. J. Numer. Meth. Fluids. 2010. Vol. 63. № 1. P. 1–21.
21. Issa R. Solution of the implicit discretized fluid flow equations by operator splitting // J. Comput. Phys. 1986. Vol. 62. № 1. P. 40–65.
22. Kraposhin M. V., Banholzer M., Pfitzner M., Marchevsky I. K. A hybrid pressure-based solver for nonideal single-phase fluid flows at all speeds // Int. J. Numer. Meth. Fluids. 2018. Vol. 88. № 2. P. 79–99.
23. United collection of hybrid Central solvers — one-phase, two-phase and multicomponent versions: [Online]. URL: https://github.com/unicfdlab/hybridCentralSolvers (Accessed: 02.07.2024).
24. Chetverushkin B. N. Kinetic schemes and quasi-gas-dynamic system of equations. CIMNE, Barcelona, Spain, 2008. 298 p.
25. Elizarova T.G. Quasi-Gas Dynamic Equations. Springer, Berlin, Heidelberg, 2009. 286 p.
26. Lyubimov A.N., Rusanov V.V. Gas Flows Past Blunt Bodies: Part II. Tables of the Gasdynamic Functions. NASA Technical Translation F-714, 1973.
27. Bondarev А. Е., Kuvshinnikov А. Е. Comparative Estimation of QGDFoam Solver Accuracy for Inviscid Flow Around a Cone // IEEE The Proceedings of the 2018 Ivannikov ISPRAS Open Conference (ISPRAS-2018). P. 82–87.