Fluids 10 00045
Fluids 10 00045
1 Computational Engineering Group, Agreste Academic Center (CAA), Federal University of Pernambuco
(UFPE), Avenida Marielle Franco, KM 59, s/n, Nova Caruaru, Caruaru 55014-900, PE, Brazil;
[email protected]
2 Laboratory of Thermal and Fluid Engineering (LATEF), School of Electrical, Mechanical and Computing
Engineering (EMC), Federal University of Goiás (UFG), Avenida Esperança, Al. Ingá, Prédio B5,
Campus Samambaia, Goiânia 74690-900, GO, Brazil; [email protected]
* Correspondence: [email protected]
† These authors contributed equally to this work.
Abstract: This study evaluates the performance of the IMERSPEC methodology combined
with the Spalart–Allmaras turbulence model for simulating fully developed turbulent flows
in a plane channel. Turbulent flows, known for their complexity, require numerical methods
that balance computational efficiency with accuracy. The IMERSPEC approach, recognized
for its spectral accuracy and efficiency, was applied alongside the Spalart–Allmaras model,
valued for its simplicity and robustness in representing turbulence, particularly in scenarios
where flow over solid surfaces is critical. Simulations were conducted at Reynolds numbers
(Reτ ) of 180, 550, and 1000, with results validated against direct numerical simulation
(DNS) data. The study investigated various grid resolutions, revealing that finer meshes
substantially enhance accuracy by mitigating velocity profile oscillations and reducing
the L2 error norm. Key findings highlight the method’s ability to accurately replicate
turbulent flow characteristics, including velocity distributions and shear stress profiles,
Academic Editor: Giuliano De Stefano while maintaining a favorable computational cost-to-accuracy ratio. This work provides
Received: 16 December 2024 valuable insights into turbulence modeling, demonstrating the potential of the IMERSPEC
Revised: 24 January 2025 methodology for practical engineering applications.
Accepted: 28 January 2025
Published: 11 February 2025 Keywords: IMERSPEC methodology; Spalart–Allmaras turbulence model; turbulent-flow
Citation: Albuquerque, L.A.V.d.; simulations
Villela, M.F.d.S.; Mariano, F.P.
Numerical Evaluation of the
IMERSPEC Methodology and
Spalart–Allmaras Turbulence Model in 1. Introduction
Fully Developed Channel Flow
Simulations. Fluids 2025, 10, 45.
Most flows encountered in nature and practical engineering applications exhibit tur-
https://doi.org/10.3390/ bulent characteristics, characterized by pressure and velocity fluctuations superimposed on
fluids10020045 the mean flow. For instance, the boundary layer in the Earth’s atmosphere is generally tur-
Copyright: © 2025 by the authors.
bulent, except under highly stable conditions, while sub-surface ocean currents also exhibit
Licensee MDPI, Basel, Switzerland. turbulence [1]. Similarly, turbulent behavior is observed in the boundary layers formed
This article is an open access article around aircraft wings, vehicles, and in the flow of natural gas and oil through pipelines [2].
distributed under the terms and Understanding turbulence and its underlying physical mechanisms necessitates a detailed
conditions of the Creative Commons
analysis of the temporal and spatial variations of these fluctuations [3,4]. Factors such as
Attribution (CC BY) license
fluid viscosity and geometry play crucial roles in shaping the diverse turbulent structures
(https://creativecommons.org/
licenses/by/4.0/).
observed in these flows [5].
they come with a high computational cost, particularly for higher Reynolds numbers.
LES offers a middle ground in terms of computational cost and accuracy, suitable for
complex geometries, whereas the SA model, being part of RANS-based methods, provides
a computationally efficient alternative for scenarios where only mean flow properties
are required.
To understand turbulent phenomena, computational fluid dynamics (CFD) relies on
advanced techniques that use high-order accuracy methods, enabling the generation of
results that closely represent the physical processes involved. Among these methods,
spectral techniques are particularly notable for their ability to combine high accuracy with
reduced computational costs compared to classical high-order methods, such as high-
order finite difference and compact schemes [12]. This efficiency gain is made possible
through the use of fast Fourier transform (FFT) [13]: while finite difference methods incur a
computational cost of O( N 2 ), where N is the number of grid points, FFT operates at a cost
of O( N log2 N ) [12]. Moreover, the projection method [14,15] enables the calculation of the
pressure field in spectral space, eliminating the need to solve the Poisson equation, one of
the most computationally expensive steps in conventional simulations.
To address complex geometries and non-periodic boundary conditions, the immersed
boundary method (IBM) [16] offers a highly effective solution. This approach introduces
a source term that acts as a body force in the Navier–Stokes equations, allowing the
representation of an immersed geometry within the flow without the need to modify the
mesh [15]. The integration of the IBM and pseudospectral Fourier methods leads to the
IMERSPEC approach [14,15]. In the physical domain, boundary conditions are applied
using the IBM, while periodic boundary conditions, which are essential for FFT calculations,
are enforced in the complementary external domain.
This study builds upon and extends the IMERSPEC methodology to simulate turbulent
flows using the RANS equations coupled with the Spalart–Allmaras turbulence model.
The simulations are validated by comparing the results with direct numerical simulation
(DNS) data for plane channel flows at various Reynolds numbers. In doing so, this work
enhances the understanding and application of IMERSPEC in capturing the complexities
of turbulent flow behavior. The research contributes to the development of efficient and
accurate computational fluid dynamics (CFD) methods, with potential applications in
aerodynamic design, fluid dynamics research, and engineering analysis.
∂ūi
= 0, (1)
∂xi
" ! #
∂ūi ∂ 1 ∂ p̄ ∂ ∂ūi ∂ū j f¯
− ui′ u′j + i ,
+ ūi ū j = − + ν + (2)
∂t ∂x j ρ ∂xi ∂x j ∂x j ∂xi ρ
where p̄ is the mean pressure [N/m2 ], ūi is the mean velocity in the i-direction [m/s],
f¯i is a source term [N/m3 ], ρ is the density [kg/m3 ], ν is the kinematic viscosity [m2 /s],
xi represents the spatial coordinate (x or y) [m], and t is time [s].
The term ui′ u′j in Equation (2), known as the Reynolds stress tensor, introduces closure
issues due to its unknown nature. To address this, Boussinesq proposed a hypothesis that
models the Reynolds stress tensor by drawing an analogy to the Stokes model for molecular
viscous stresses [5]. The resulting model is given by Equation (3):
Fluids 2025, 10, 45 4 of 17
!
∂ūi ∂ū j 2
−ui′ u′j = νt + + kδij , (3)
∂x j ∂xi 3
where k = 21 ui′ ui′ represents the turbulent kinetic energy, δij is the Kronecker delta, and νt is
the turbulent kinematic viscosity. Unlike molecular viscosity, an inherent fluid property,
turbulent viscosity is a flow-dependent characteristic.
The turbulent viscosity νt can be determined using turbulence models such as Spalart–
Allmaras, employed in this study. Additionally, the turbulent kinetic energy contributes
to a modified pressure term, p∗ , which combines the mean pressure and the turbulence
effects. Considering these modifications, Equation (2) can be rewritten as:
" !#
∂ūi ∂ 1 ∂ p̄∗ ∂ ∂ūi ∂ū j f¯
+ i.
+ ūi ū j = − + (ν + νt ) + (4)
∂t ∂x j ρ ∂xi ∂x j ∂x j ∂xi ρ
Here, ⃗x denotes the position of a fluid particle, while ⃗ X represents the position of a
particle adjacent to the solid interface, as illustrated in Figure 1. The source term assumes
non-zero values exclusively at points coinciding with the immersed geometry, enabling the
Eulerian field to detect the presence of the solid interface [14].
equations, applied to the fluid particles residing at the fluid–solid boundary, as expressed
in Equation (6). The notations and variables used are consistent with Equation (4), but the
calculations are restricted to the Lagrangian interface (Γ) shown in Figure 1.
" !#
⃗ ∂Ūi ⃗ ∂
⃗ 1 ∂ P¯∗ ⃗ ∂ ∂Ūi ∂Ūj
Fi ( X, t) = ( X, t) + Ū Ū ( X, t) + ( X, t) − (ν + νt ) + . (6)
∂t ∂X j i j ρ ∂Xi ∂X j ∂X j ∂Xi
By employing the temporal parameter U ∗ introduced by [17] and discretizing the time
operator, Equation (7) is obtained. It is important to note that the explicit Euler method
is utilized for the time derivative in this equation, primarily for illustrative and didactic
purposes. Nevertheless, in the present study, a more accurate fourth-order Runge–Kutta
method is applied for time discretization [18].
X, t + ∆t) − Ui∗ ( ⃗
Ui ( ⃗ X, t) + Ui∗ ( ⃗
X, t) − Ui ( ⃗
X, t)
Fi ( ⃗
X, t) = + RHSi ( ⃗X, t). (7)
∆t
where ∆t is the timestep, and
" !#
⃗ ∂
⃗ 1 ∂P∗ ⃗ ∂ ∂Ūi ∂Ūj
RHSi ( X, t) = Ū Ū ( X, t) + ( X, t) − (ν + νt ) + . (8)
∂X j i j ρ ∂Xi ∂X j ∂X j ∂Xi
Ui∗ ( ⃗
X, t) − Ui ( ⃗
X, t)
+ RHSi ( ⃗X, t) = 0, (9)
∆t
X, t + ∆t) − Ui∗ ( ⃗
Ui ( ⃗ X, t)
Fi ( ⃗
X, t) = . (10)
∆t
X, t + ∆t) = UFI is the velocity at the immersed boundary, and Ui∗ ( ⃗
Here, Ui ( ⃗ X, t) is
expressed as:
u∗ ( ⃗ x=⃗
i X, t ) if ⃗ X,
Ui ( ⃗
X, t) = (11)
0 if ⃗x ̸= ⃗
X.
Equation (9) is solved in the Eulerian domain using the Fourier spectral space. The
intermediate velocity ui∗ (⃗x, t) is interpolated onto the Lagrangian domain, resulting in
Ui∗ ( ⃗
X, t) as per Equation (10). This value is subsequently mapped back to the Eulerian
collocation points, and the final velocities in the Eulerian domain are updated using
Equation (12).
ui (⃗x, t + ∆t) = ui∗ (⃗x, t) + ∆t f i . (12)
where k j and ub̄j denote the wave vector and the Fourier-transformed velocity vector,
respectively, and νe f = ν + νt .
Fluids 2025, 10, 45 6 of 17
Equation (13) establishes the orthogonality between the wave number vector, k i , and
the transformed velocity, ub̄i (⃗k, t). This relationship defines a divergence-free vector plane,
referred to as the π-plane, which is orthogonal to the wave number vector and contains the
transformed velocity [5]. Analyzing the components of Equation (14) reveals distinct spatial
orientations: the transient and viscous terms lie within the π-plane, while the pressure
gradient term is perpendicular to it. The orientation of the nonlinear term, however, remains
initially undetermined. By combining all components of Equation (14), the equation is
resolved as follows:
∂(\
ūi ū j ) \ ∂ū !
∂ub̄i ∂ ∂ ū i j f
b̄
+ i = 0.
+ p∗ −
+ ik i c̄ ν + (15)
∂t ∂x j ∂x j e f ∂x j ∂xi ρ
| {z }
∈π | {z }
∈π
where ℘im is the projection tensor. The pressure gradient term lies orthogonal to the
π-plane, decoupling the pressure and velocity fields in Fourier space. Nevertheless, the
pressure field can be reconstructed via post-processing, as detailed by [15].
The nonlinear term introduces a product of transformed functions which, based on the
properties of Fourier transforms, corresponds to a convolution operation. This operation is
represented as a convolution integral, whose computational solution is expensive compared
to other methods. To reduce this cost, the pseudospectral Fourier method is utilized. In
this approach, the momentum equation in Fourier space, formulated using the projection
method, is expressed as:
\∂ū !
∂ub̄i (⃗k, t) ∂ūm
Z Z
j
= −ik j ℘im m (⃗r ) u
uc̄ b̄j (⃗k −⃗r ) d⃗r + ik j ℘im νc (⃗r ) + (⃗k −⃗r ) d⃗r. (17)
∂t ⃗k=⃗r +⃗s ⃗k=⃗r +⃗s e f ∂x j ∂xm
The nonlinear term can be treated in various forms. Among these, the skew–symmetric
formulation is preferred due to its superior stability and accuracy [19]. In this study, the
pseudospectral method is employed to solve the nonlinear term, wherein the velocity prod-
uct is computed in physical space and subsequently transformed back into spectral space.
The terms on the right-hand side of Equation (18) correspond, respectively, to the
production, destruction, molecular and turbulent diffusion, and dissipation of viscosity.
In this context, d denotes the distance to the nearest wall. The turbulent viscosity νt is
Fluids 2025, 10, 45 7 of 17
calculated using the modified viscosity ν̃, incorporating a damping function f ν1 to account
for near-wall effects:
νt = ν̃ f ν1 , (19)
where
χ3
f ν1 = , (20)
χ3 + Cν31
and
ν̃
χ= . (21)
ν
Far from the wall, the damping function f ν1 asymptotically approaches unity, reducing
νt to ν̃. The production term depends on the distance from the wall and is modulated by
the function f ν2 , which is defined as:
ν̃
S̃ = S + fν , (22)
κ 2 d2 2
where
χ
f ν2 = 1 − . (23)
1 + χ f ν1
In Equation (22), the mean rate of strain S is defined in terms of the vorticity tensor Ω:
q
S= 2Ωij Ωij , (24)
where !
1 ∂ui ∂u j
Ωij = − . (25)
2 ∂x j ∂xi
The wall function f w , which plays a crucial role near the boundary layer, amplifies the
production term in proximity to the wall and diminishes in magnitude as the distance from
the wall increases. It is defined as:
!1
6
Cw3 6
f w = g 1 + 6 6
, (26)
g + Cw3
where
g = r + Cw2 (r6 − r ), (27)
and
ν̃
r= . (28)
S̃κ 2 d2
The model relies on empirical constants, summarized in Table 1, which are calibrated
based on well-known flows, especially turbulent boundary layers:
Constant Value
Cb1 1+Cb2
Cw1 +
κ2 σ
Cw2 0.3
Cw3 2
κ 0.41
Cv1 7.1
σ 2
3
Fluids 2025, 10, 45 8 of 17
Table 1. Cont.
Constant Value
Cb1 0.1355
Cb2 0.622
Ct1 1
Ct2 2
Ct3 1.1
Ct4 2
In this study, the Spalart–Allmaras equation (Equation (18)) is solved using the Fourier
pseudospectral method, necessitating a transformation of the equation into spectral space.
Fourier transform is applied to the dependent variables, thereby converting the spatial
derivatives into multiplications by complex numbers corresponding to the frequencies in
spectral space.
The time derivative of ν̃ in physical space is directly transformed as:
∂ν̃ ∂ν̃ˆ
→ . (29)
∂t ∂t
The convective term is transformed by applying the derivative property of Fourier
transform:
∂(u j ν̃)
→ ik j uc
j ν̃. (30)
∂x j
The production and destruction terms are transformed into spectral space as follows:
2 \ 2
ν̃ ν̃
−Cw1 f w → −Cw1 f w . (32)
d d
The viscous diffusion term is transformed into spectral space by considering the
application of the derivative and convolution in spectral space:
" #
1 ∂ ∂ν̃ 1
(ν + ν̃) ν + ν̃) ∗ ik j ν̃ˆ .
→ ik j (\ (33)
σ ∂x j ∂x j σ
Finally, the turbulent diffusion term is transformed into Fourier spectral space as:
!
cb2 ∂ν̃ ∂ν̃ cb2 2 ˆ 2
→ k ν̃ . (34)
σ ∂x j ∂x j σ j
\ 2
∂ν̃ˆ ν̃ 1 \ ˆ
cb2 2 2
+ ik j uc
j ν̃ = c c̃ν̃ − C f
S w1 w − ik ( ν + ν̃ )∗ ik ν̃ + k ν̃ˆ . (35)
∂t b1
d σ j j
σ j
• Non-divergent form: In the non-divergent form, the terms are expressed directly in
terms of the derivative of the product of the functions:
∂ν̃ \
ū j = ū j ik j ν̃. (37)
∂x j
1 h [ \i
j ν̃ =
ik j uc̄ ik (ū ν̃) + ū j ik j ν̃ . (38)
2 j j
This skew–symmetric treatment preserves the balance of the essential physical proper-
ties of the flow, while providing enhanced numerical stability.
The imposition of boundary conditions for the modified turbulent viscosity ν̃ is
achieved using the immersed boundary method, which treats the boundary conditions
as additional force sources within the computational domain. This is accomplished by
introducing a virtual force that acts on the variable ν̃, guiding it to satisfy the imposed
boundary condition. In the case of the no-slip condition, the force is applied in such a way
that the tangential velocity at the wall is zero.
In the Spalart–Allmaras equation, the imposition of a no-slip boundary condition can
be mathematically represented by modifying the convective term in the transport equation
for ν̃, as follows:
2 " # !
∂ν̃ ∂(ū j ν̃) ν̃ 1 ∂ ∂ν̃ cb2 ∂ν̃ ∂ν̃
+ = cb1 S̃ν̃ − Cw1 f w + (ν + ν̃) + + f bc , (39)
∂t ∂x j d σ ∂x j ∂x j σ ∂x j ∂x j
where fbc represents the virtual force imposed to satisfy the no-slip condition. This force is
calculated iteratively to ensure that ν̃ reaches the desired value at the boundary:
where λ is a control parameter that adjusts the intensity of the applied force, and ν̃bc is the
reference value of ν̃ at the boundary, which is zero for the no-slip condition.
The MDF is used to iteratively adjust the virtual force in order to correct any deviation
from the desired boundary condition. Each successive application of the force brings ν̃
closer to the reference value at the boundary, thus ensuring the exact imposition of the
no-slip condition.
3. Results
This study investigates fully developed, incompressible, statistically steady, and homo-
geneous turbulent flow in the axial direction (x). The predictions are validated against direct
numerical simulation (DNS) data from [21,22], which provided a detailed examination of
Fluids 2025, 10, 45 10 of 17
this flow. The Reynolds numbers (Reτ ), defined based on the friction velocity (uτ ), were set
at 180, 950, and 1000. According to [23], a value of Reτ ∼ 180 is commonly estimated as the
transition point between laminar and turbulent flow in channel configurations. Figure 2
depicts the geometry and coordinate system used in the analysis.
The flow is modeled as incompressible, with no heat transfer, within a plane channel of
axial length L x = 10 m and width Ly = 1 m. It is treated as two-dimensional on average, ne-
glecting edge effects in the direction normal to the study plane. This test case is particularly
appropriate because the Spalart–Allmaras turbulence model was specifically calibrated to
capture the characteristics of channel flows. Therefore, a robust implementation of the SA
model using the IMERSPEC methodology is expected to produce results closely aligned
with DNS data.
To assess the computational performance and predictive accuracy of the model in
two-dimensional simulations, a uniform grid of 64 × 128 cells was employed. At the walls,
a no-slip condition was imposed for the velocities u and vs., while periodic boundary
conditions were applied in the axial direction (x). The solution was computed under a
prescribed mean pressure gradient (∆p = 0) to achieve the desired Reynolds number. The
initial condition for the simulation assumed a uniform flow across the entire domain.
1
The kinematic viscosity was defined as ν = Re , and the inflow velocity U∞ was
set to 1.0 m/s. Initially, the stream-wise velocity component u was uniformly set to the
inflow velocity throughout the domain, while the vertical velocity component vs. was
initialized to zero. At the upper and lower walls, the no-slip condition ensured that u = 0.
This condition was enforced using the immersed boundary method at each time step.
To maintain a Courant–Friedrichs–Lewy (CFL) number of approximately 0.1 across the
domain, a time step of ∆t = 2 × 10−4 s was chosen, corresponding to a dimensionless time
step ∆t∗ = 2.8 × 10−2 ; this was based on the mean velocity and half the channel height (H).
From the initial velocity field, the governing equations were integrated in time until
the numerical solution reached a statistically steady state. The attainment of this state was
determined by monitoring the convergence behavior of key parameters, such as the wall
shear stress and the mean velocity at the channel center line (y = H). Statistical steady-state
conditions were confirmed when these parameters displayed negligible variations over
time, exhibiting approximately periodic behavior.
Figure 3 presents the normalized mean velocity profile as a function of the normalized
distance across the channel width for fully developed flow at Reτ = 1000, compared to
DNS data. Similarly, Figure 4 shows the viscous shear stress profile for Reτ = 1000, also
compared with DNS data. For Reτ = 550, the corresponding mean velocity and shear stress
profiles are displayed in Figures 5 and 6, respectively. For Reτ = 180, the mean velocity
profile is shown in Figure 7, while Figure 8 presents the viscous shear stress profile.
Fluids 2025, 10, 45 11 of 17
Figure 3. Normalized mean velocity profile for Reτ = 1000, compared with the results of [21].
Figure 4. Viscous shear stress profile for Reτ = 1000, compared with the results of [21].
Figure 5. Normalized mean velocity profile for Reτ = 550, compared with the results of [21].
Fluids 2025, 10, 45 12 of 17
Figure 6. Viscous shear stress profile for Reτ = 550, compared with the results of [21].
Figure 7. Normalized mean velocity profile for Reτ = 180, compared with the results of [22].
Figure 8. Viscous shear stress profile for Reτ = 180, compared with the results of [22].
As the flow reaches a fully turbulent state, the agreement with DNS data improves.
With an increase in the Reynolds number, the flow becomes progressively more turbulent,
leading to higher turbulent stresses. Additionally, the mean velocity profile becomes more
uniform across the channel height due to enhanced momentum transfer in the y-direction.
Fluids 2025, 10, 45 13 of 17
For Reτ = 1000, simulations were conducted using different mesh configurations to
evaluate the impact of spatial resolution on the accuracy of the results. Figures 9–12 show
the mean velocity profiles obtained for meshes with resolutions of 64 × 32, 64 × 64, 64 × 128,
and 64 × 256 grid points.
Figure 10. Mean velocity profile for Reτ = 1000 with a 64 × 64 mesh.
Fluids 2025, 10, 45 14 of 17
Figure 11. Mean velocity profile for Reτ = 1000 with a 64 × 128 mesh.
Figure 12. Mean velocity profile for Reτ = 1000 with a 64 × 256 mesh.
The results presented in Figures 9–11 reveal the presence of oscillations in the velocity
profiles, particularly in coarser meshes such as 64 × 32 and 64 × 64. These oscillations are
characteristic of spectral methods, which tend to be highly sensitive to mesh resolution,
especially in high-Reynolds-number turbulent-flow simulations. This observation under-
scores the necessity for finer meshes to better capture the dynamic behavior associated
with turbulence.
For the 64 × 128 mesh, the L2 norm, which quantifies the difference between the
numerical results and the reference DNS data, was found to be 1.992 × 10−2 . Using the
Fluids 2025, 10, 45 15 of 17
finer 64 × 256 mesh, this error was significantly reduced to L2 = 5.176 × 10−3 , highlighting
the substantial improvement in accuracy with increased mesh resolution. These results
emphasize the crucial role of mesh refinement in obtaining numerical solutions that better
approximate the physical reality represented by DNS data. For reference, the mesh used in
the DNS simulation was n x = 10,240 and nz = 7680 [21].
Figure 13 illustrates the velocity field for Reτ = 1000, showcasing the typical behavior
of a fully developed turbulent flow. In this flow regime, the transverse velocity gradient
reflects the intense momentum transfer in the direction normal to the main flow, resulting
in a flatter velocity profile across the channel cross-section. This behavior aligns with the
expected dynamics for high-Reynolds-number flows in channels, where turbulent stresses
play a dominant role in the redistribution of momentum.
4. Discussion
The results underscore the effectiveness of the Spalart–Allmaras turbulence model,
implemented using the IMERSPEC methodology, in accurately capturing the key character-
istics of turbulent flow in plane channels. Validation against direct numerical simulation
(DNS) data from the literature [21,22] demonstrated strong agreement, particularly for
simulations conducted with refined mesh configurations.
Mesh refinement tests at Reτ = 1000 highlighted the sensitivity of the spectral method
to spatial resolution. For coarser meshes (64 × 32 and 64 × 64), oscillations appeared in the
mean velocity profiles, indicating that, while the spectral method provides high accuracy
and convergence efficiency, achieving reliable results requires sufficient spatial refinement
to resolve the intricate dynamics of turbulent flows. Conversely, finer meshes (64 × 128
and 64 × 256) yielded significantly improved agreement with DNS data, as evidenced by a
marked reduction in the L2 norm error. This improvement demonstrates the importance
of mesh refinement in accurately capturing momentum exchange and velocity gradients
characteristic of turbulent regimes.
The mean velocity profiles and viscous shear stresses confirmed the expected behavior
of fully developed turbulent flows. At higher Reynolds numbers, turbulence intensity
increases, resulting in greater momentum redistribution in the transverse direction and
flatter velocity profiles. These findings align with previous DNS studies, which emphasize
the predominance of turbulent stresses in the dynamics of high-Reynolds-number flows.
The results further reinforce the suitability of the Spalart–Allmaras model in scenarios
where flow over solid boundaries and high Reynolds numbers are critical, making it a
robust choice for practical engineering applications.
Fluids 2025, 10, 45 16 of 17
5. Conclusions
This study demonstrated that the IMERSPEC methodology, combined with the Spalart–
Allmaras turbulence model, is a highly promising approach for simulating turbulent flows
in simplified geometries, particularly for applications that demand a balance between
computational efficiency and accuracy. The results highlight the method’s ability to accu-
rately capture the key characteristics of turbulent flows while maintaining manageable
computational costs.
Future research should focus on extending this methodology to more complex and
realistic geometries, such as flows around cylinders and airfoils. Additionally, the devel-
opment and implementation of advanced mesh refinement strategies could enhance the
computational efficiency of the spectral method without sacrificing accuracy. Investigating
the integration of other turbulence models with the IMERSPEC methodology also holds
significant potential, as it would enable the exploration of a broader range of turbulent
scales and flow regimes. Moreover, extending the approach to three-dimensional simula-
tions represents a critical next step for achieving a more comprehensive understanding and
modeling of turbulent phenomena.
The findings of this work not only affirm the effectiveness of combining the Spalart–
Allmaras model with the IMERSPEC methodology but also lay a strong foundation for
future studies involving more complex flow scenarios. This contributes meaningfully to
the ongoing advancement of computational fluid dynamics, particularly in the simulation
of turbulent flows.
Funding: This research was funded by Eletrobras and the Research and Technological Development
Program (P&D) of ANEEL, under grant number ANEEL PD-00394-1906/2019.
Data Availability Statement: Data can be accessed upon request to any author of this manuscript.
Acknowledgments: The authors gratefully acknowledge the support of Eletrobras and the ANEEL
Research and Technological Development Program (P&D) in the development of this work.
References
1. Ruan, X. Note on the Bulk Estimate of the Energy Dissipation Rate in the Oceanic Bottom Boundary Layer. Fluids 2022, 7, 82.
[CrossRef]
2. Tennekes, H.; Lumley, J.L. A First Course in Turbulence; The MIT Press: Cambridge, MA, USA, 1972.
3. Abramov, R.V. Turbulence via Intermolecular Potential: Viscosity and Transition Range of the Reynolds Number. Fluids 2023,
8, 101. [CrossRef]
4. Sofiadis, G.; Sarris, I. Turbulence Intensity Modulation by Micropolar Fluids. Fluids 2021, 6, 195. [CrossRef]
5. Silveira Neto, A. Escoamentos Turbulentos—Análise Física e Modelagem Teórica; Composer: Uberlândia, Brazil, 2020.
6. Pope, S.B. Turbulent Flows; Cambridge University Press: Cambridge, UK, 2000.
7. Kadivar, M.; Tormey, D.; McGranaghan, G. A review on turbulent flow over rough surfaces: Fundamentals and theories. Int. J.
Thermofluids 2021, 10, 100077. [CrossRef]
Fluids 2025, 10, 45 17 of 17
8. Magalhães, G.M. Modelagem Matemática e Computacional de Escoamentos Turbulentos Multifásicos em Malha Adaptativa
Bloco Estruturada Utilizando Método Multi Direct Forcing. Master’s Thesis, Universidade Federal de Uberlândia, Uberlândia,
Brazil, 2022.
9. Versteeg, H.K.; Malalasekera, W. An Introduction to Computational Fluid Dynamics: The Finite Volume Method; Pearson Education:
Harlow, UK, 2007.
10. D’Alessandro, V.; Montelpare, S.; Ricci, R.; Zoppi, A. Numerical modeling of the flow over wind turbine airfoils by means of
Spalart–Allmaras local correlation based transition model. Energy 2017, 130, 402–419. [CrossRef]
11. D’Alessandro, V.; Montelpare, S.; Ricci, R. Assessment of a Spalart–Allmaras Model Coupled with Local Correlation Based
Transition Approaches for Wind Turbine Airfoils. Appl. Sci. 2021, 11, 1872. [CrossRef]
12. Canuto, C.; Hussaini, M.Y.; Quarteroni, A.; Zang, T.A. Spectral Methods: Evolution to Complex Geometries and Applications to Fluid
Dynamics, 1st ed.; Springer: Berlin, Germany, 2007.
13. Cooley, J.W.; Tukey, J.W. An algorithm for the machine calculation of complex Fourier series. Math. Comput. 1965, 19, 297–301.
[CrossRef]
14. Nascimento, A.A.; Mariano, F.P.; Padilla, E.L.M.; Silveira-Neto, A. Comparison of the convergence rates between Fourier
pseudospectral and finite volume method using Taylor-Green vortex problem. J. Braz. Soc. Mech. Sci. Eng. 2020, 42, 491.
[CrossRef]
15. Mariano, F.P.; de Queiroz Moreira, L.; Nascimento, A.A.; Silveira-Neto, A. An improved immersed boundary method by coupling
of the multi-direct forcing and Fourier pseudospectral methods. J. Braz. Soc. Mech. Sci. Eng. 2022, 44, 388. [CrossRef]
16. Peskin, C.S. Flow patterns around heart valves: A numerical method. J. Comput. Phys. 1972, 10, 252–271. [CrossRef]
17. Wang, J.; Fan, J.; Luo, K. Combined multi-direct forcing and immersed boundary method for simulating flows with moving
particles. Int. J. Multiphase Flow 2008, 34, 283–302. [CrossRef]
18. Allampalli, V.; Hixon, R.; Nallasamy, M.; Sawyer, S.D. High-accuracy large-step explicit Runge–Kutta (HALE-RK) schemes for
computational aeroacoustics. J. Comput. Phys. 2009, 228, 3837–3850. [CrossRef]
19. Albuquerque, L.A.V.d.; Villela, M.F.d.S.; Mariano, F.P. Numerical Simulation of Flows Using the Fourier Pseudospectral Method
and the Immersed Boundary Method. Axioms 2024, 13, 228. [CrossRef]
20. Canuto, C.; Hussaini, M.Y.; Quarteroni, A.; Zang, T.A. Spectral Methods: Fundamentals in Single Domains; Springer: New York, NY,
USA, 2006.
21. Lee, M.; Moser, R.D. Direct numerical simulation of turbulent channel flow up to Reτ = 5200. J. Fluid Mech. 2015, 774, 395–415.
[CrossRef]
22. del Álamo, J.C.; Jiménez, J. Spectra of the very large anisotropic scales in turbulent channels. Phys. Fluids 2003, 15, L41–L44.
[CrossRef]
23. Schlichting, H.; Kestin, J. Boundary-Layer Theory, 8th ed.; Springer: Berlin/Heidelberg, Germany, 2017.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual
author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to
people or property resulting from any ideas, methods, instructions or products referred to in the content.