3.1. Reference Case
For validation purposes,
Figure 3 shows the predicted temperature contours obtained through direct numerical simulation (DNS) using the Rayleigh numbers (Ra) reported by Trias et al. [
33]. It is important to note that these authors also used DNS and the finite volume method (FVM) to model natural convection in a cavity with an aspect ratio of A = 4, achieving excellent agreement between the two studies. The temperature contours reveal significant agreement with those documented in the reference work. In both scenarios, the characteristic thermal structures of natural convection in closed cavities are evident, demonstrating complex convective circulation driven by the thermal gradient between the hot and cold walls. For Ra = 6.4 × 10
8, a relatively stable convective pattern is observed, characterized by well-defined thermal layers and smooth transitions between high- and low-temperature regions, indicating that turbulence has not yet entirely dominated the fluid flow.
As Ra increases to 2 × 10
9, the thermal layers start to deform due to stronger convective effects, leading to more irregularities in the isotherms, which indicates the beginning of the transition to turbulence. The comparison with the results reported by Trias et al. [
33] shows a similar trend, confirming that direct numerical simulation (DNS) can effectively capture these phenomena. At Ra = 10
10, the flow becomes fully turbulent, as shown by the more irregular isotherms and increased thermal mixing within the cavity, along with the development of extremely fine thermal boundary layers along the hot and cold walls—a typical feature of high-Ra flows. Compared to the results of Trias et al. [
33], the overall structure of the temperature contours remains consistent, although slight differences may exist due to mesh refinement and the discretization methods used in both studies. In conclusion, the comparison of the temperature contours obtained in this study with those reported by Trias et al. [
33] confirms the validity of the direct numerical simulation used here, clearly illustrating the evolution of the thermal flow structure as Ra increases—from more stable configurations at Ra = 6.4 × 10
8 to fully turbulent flow at Ra = 1 × 10
10. This indicates that the mesh and numerical methods employed are suitable for modeling natural convection effects at high Rayleigh numbers, increasing confidence in the accuracy of the results [
34].
In addition to the qualitative comparison of thermal contours, a quantitative evaluation of the results was conducted using the average Nusselt number calculated along the hot wall.
Table 3 presents the values obtained in this study and those reported by Trias et al. [
33] for similar Rayleigh numbers. For Ra = 6.4 × 10
8, the Nusselt number calculated in this work was 49.98, while Trias et al. [
33] reported a value of 49.23, resulting in a relative difference of only 1.52%. For Ra = 2 × 10
9, the value obtained was 65.78, compared to 66.19 reported by Trias et al. [
33], representing a difference of 0.62%. Finally, for Ra = 1 × 10
10, the Nusselt number was 98.2, whereas Trias et al. [
33] reported 100.6, yielding a difference of 2.39%. These minor discrepancies are within the expected range. These differences can be attributed to variations in mesh refinement, numerical schemes, and convergence criteria, all of which can have a slight influence on the prediction of thermal flow in the turbulent regime. Nonetheless, the excellent agreement between both studies—both in the temperature contours and in the calculated Nusselt numbers—confirms the validity and accuracy of the direct numerical simulation (DNS) employed in this work, as well as the model’s ability to reliably reproduce the thermal behavior of natural convection at high Rayleigh numbers.
Building on the modeling approach previously described, this study was conducted under steady-state assumptions, utilizing spatially averaged values obtained from fully converged numerical results. In contrast, Trias et al. [
33] performed unsteady simulations, averaging results over time once a statistically steady regime was reached. In these cases, flow quantities oscillate around stable mean values, and the Nusselt number is derived from temporal integration. Despite the differences in averaging methods, the close match in both thermal distributions and average Nusselt numbers confirms the accuracy of this numerical strategy and its effectiveness in capturing natural convection behavior at high Rayleigh numbers.
3.2. Nusselt Number
From an applied perspective, one of the most relevant parameters for quantifying heat transfer efficiency in convection systems is the Nusselt number (Nu). This dimensionless number measures the additional heat transferred due to fluid motion (natural convection) compared to the heat transferred by conduction alone if the fluid were at rest. In the context of differentially heated cavities, where the fluid remains confined and the lateral walls are maintained at different temperatures, the Nusselt number provides a direct measure of the intensity of convective circulation induced by temperature gradients and is determined as follows.
To evaluate the accuracy of the temperature profiles calculated using direct numerical simulation (DNS) and those obtained with the κ-ε turbulence model, the percentage difference between the average Nusselt numbers was computed as a function of the Rayleigh number. As shown in
Table 4, for Ra = 10
3 and Ra = 10
4, within the laminar regime, the percentage differences are moderate, with values of 5.17% and 2.45%, respectively. However, these differences increase significantly to 17.64% for Ra = 10
5 and 19.39% for Ra = 10
6. This trend is attributed to the inability of the κ-ε model to adequately characterize the dominant diffusive effects in the thermal boundary layers, whose resolution is critical in flows with strong conduction–convection interaction, as occurs within this Rayleigh number range [
3,
4,
8,
9,
18]. This limitation arises from the isotropic nature of the turbulent viscosity formulation in the κ-ε model, which fails to capture the intense thermal gradients and the anisotropic structures that develop within the boundary layer, particularly near the hot and cold walls, where heat transport is highly localized. In contrast, DNS directly resolves the flow dynamics without needing closure models, allowing for a more accurate representation of boundary layer thinning and its impact on the Nusselt number. As the Rayleigh number increases beyond Ra = 10
6, the flow becomes fully turbulent, and the κ-ε model behaves more consistently. For Ra = 10
7 and Ra = 10
8, the percentage difference decreases to 17.70% and 12.98%, respectively, indicating a gradual improvement in the prediction of heat transfer. This improvement can be attributed to the fact that, within this range, turbulent effects begin to dominate over conductive mechanisms, making the κ-ε model more representative of the actual flow physics.
Finally, for Ra = 109 and Ra = 1010, excellent agreement is observed between both approaches, with reduced errors of 5.61% and 1.16%, respectively. These results indicate that the κ-ε model can adequately capture the global flow behavior under fully developed turbulent conditions, where convective transport predominates and the thermal boundary layers, although thin, exhibit dynamics better represented by the turbulence model formulation.
Table 5 compares the average Nusselt numbers reported by various authors with those calculated in this study as a function of the Rayleigh number, which ranges from Ra = 10
3 to Ra = 10
10. In addition to these results,
Table 5 includes the numerical methods used by each author, enabling an analysis of the differences between these approaches. The table shows that the values obtained in this study demonstrate excellent agreement with reference results from the literature, particularly those reported by Dixit and Babu [
13] and Goloviznin et al. [
8], who utilized high-accuracy schemes to solve natural convection in two-dimensional square cavities. The range 10
3 ≤ Ra ≤ 10
8, the Nusselt numbers calculated in this work closely aligned with previous studies, validating the precise resolution of thermal boundary layers and the appropriate treatment of boundary conditions in the current model.
Additionally, for validation purposes, the natural convection problem described in this study was also solved using the orthogonal collocation method with Legendre polynomials, employing the computational code NEWCOL2L (written in FORTRAN 90) with a 71 × 71 node mesh, achieving Rayleigh numbers up to 10
9 through DNS. The discretized equations were solved using the modified Newton method [
35] with LU factorization, and the partial derivatives of the Jacobian matrix were approximated using finite differences. A detailed description of the procedure can be found in Molina-Herrera et al. [
2].
Table 3 compares the average Nusselt numbers obtained using the orthogonal collocation method with those calculated in this study, demonstrating excellent agreement. This validates the mesh employed, as the results demonstrate remarkable consistency regardless of the numerical method used.
It is important to note that experimental data for validation were not available in this numerical study.
Table 5 presents a comparative analysis of the average Nusselt numbers obtained through DNS in the current work alongside those reported in the literature by various authors using different numerical methods. This comparison includes values collected via finite volume, finite element, and spectral methods across a wide range of Rayleigh numbers. Additionally, the present DNS results were verified using the orthogonal collocation method implemented in Fortran through the NEWCOL2L code [
35], utilizing Legendre Polynomials with a 71 × 71 mesh. Although these results are not experimental, they provide a reliable verification framework that supports the accuracy and consistency of the numerical solutions presented in this study, particularly at high Rayleigh numbers where boundary layer resolution is critical.
3.3. Macroscopic Energy Balance
Another criterion for validating the accuracy of the numerical method involves calculating the percentage error in the macroscopic energy balance for a two-dimensional square cavity with adiabatic top and bottom walls, as illustrated in
Figure 4. Under these conditions, the macroscopic energy balance confirms that the heat transferred by convection from the hot wall into the fluid must equal the heat leaving the cavity through the cold wall [
2]. This comparison is made by quantifying the net heat flux at the walls using Expressions (40)–(42).
The heat flow in the horizontal walls should be zero. As the solution is numerical, the expected values should be minimal concerning the Nusselt number in question. Therefore, it is also necessary to evaluate them to ensure that the calculated Nusselt number is accurate.
Table 6 presents the macroscopic energy balance calculated using direct numerical simulation (DNS) and the κ-ε turbulence model. In the DNS results, the balance remains effectively zero up to intermediate Rayleigh numbers (Ra ≤ 10
6), with slight increases of 0.0001, 0.0004, 0.0003, and 0.0018 observed at higher Rayleigh numbers (10
7 ≤ Ra ≤ 10
10), which confirms the global conservation of energy, and the precision of the mesh employed. In contrast, the κ-ε model maintains a zero balance in laminar regimes, exhibiting moderate increases of 0.0002 and 0.0003 during transitions to higher Rayleigh numbers, with values that match those obtained by DNS in the most demanding cases. This consistency indicates that direct numerical simulation can accurately capture both laminar and turbulent regimes. At the same time, the κ-ε model provides comparable results, thereby validating the numerical strategy and mesh design used in this study. Moreover, the κ-ε model exhibits a zero balance in laminar regimes (Ra ≤ 10
4), with moderate increases of 0.0002 and 0.0003 as the flow transitions to higher Rayleigh numbers (Ra ≥ 10
5), reaching values identical to those of DNS at Ra = 10
9 and 10
10. This agreement further supports the hypothesis that DNS can precisely capture both laminar and turbulent regimes. At the same time, the κ-ε model delivers comparable outcomes, confirming the effectiveness of the numerical approach and meshing strategy employed.
Nevertheless, the global errors in the κ-ε model remain at small values. Its more stable performance under high Rayleigh conditions can be explained by the model’s average nature, which filters out high-frequency fluctuations and prioritizes macroscopic energy conservation, albeit at the cost of reduced local accuracy in the velocity and temperature profiles. This observation aligns with findings reported by Wilcox [
11], who noted that RANS models tend to exhibit a good global energy balance [
16,
32].
As in the validation of the Nusselt number, the NEWCOL2L code was also used to evaluate the percentage errors in the macroscopic energy balance using direct numerical simulation (DNS).
Table 5 shows that these errors are minor and tend to approach zero within the range of 10
3 ≤ Ra ≤ 10
10, validating global energy conservation within the cavity. This behavior is attributed to the orthogonal collocation method, which discretizes the differential operators from boundary to boundary, resulting in high accuracy [
25]. Under these conditions, the energy balance error is practically negligible, although CPU time remains significant. While the percentage errors obtained with DNS and the κ-ε turbulence model stay within an acceptable range, the results underscore the importance of finer mesh refinement in boundary regions, particularly where intense thermal gradients occur. However, due to current computational limitations, achieving this level of refinement may be impractical for simulations at higher Rayleigh numbers.
Table 7 compares the percentage errors in the macroscopic energy balance calculated using direct numerical simulation (DNS) across various mesh configurations, with the number of boundary layers ranging from 6 to 10, a stretching factor of 2, a total layer thickness of 0.01 m, and an Extremely Fine mesh. This comparison enables the assessment of the impact of local refinement in critical regions of the domain, particularly along the hot and cold walls, where significant thermal gradients are present. The results indicate that while the reduction in error is notable when increasing from 6 to 8 boundary layers, the difference between 8 and 10 layers becomes minimal. This suggests that beyond 8 layers, the model sufficiently captures boundary layer dynamics without requiring further refinement, which would otherwise increase computational costs. Therefore, the configuration with 8 boundary layers is optimal, providing an efficient balance between numerical accuracy and computational demand. Furthermore, for Ra = 10
10, for BL = 8 and BL = 10, the error in the overall energy balance is practically 0.00183%. However, the CPU time is smaller for BL = 8, which confirms the effectiveness of the mesh optimization.
3.4. Comparison of Temperature and Velocity Profiles Between DNS and the κ-ε Model
Continuing the analysis of the differentially heated square cavity,
Figure 5 shows the evolution of the dimensionless temperature profiles along the cavity length (X), comparing results from direct numerical simulation (solid lines) with those from the κ-ε turbulence model (dash-dotted lines). For Ra = 10
3, the flow is fully conductive, as indicated by the straight green lines, with a 2.59% difference between the two models. This difference grows to 5.14% for Ra = 10
4 and 5.8% for Ra = 10
5. The most significant difference occurs at Ra = 10
6, reaching 12.5%, indicating that the κ-ε model underestimates the temperature near the walls due to turbulent viscosity, which reduces the conductive contribution in the thermal boundary layer of the energy equation. This behavior aligns with the findings of Barakos et al. [
3], who reported that the κ-ε turbulence model accurately reproduces flow characteristics at high Ra but exhibits deviations near the walls in laminar or transitional regimes. Starting with Ra = 10
7, the difference between the models decreases sharply, and for Ra = 10
8, only a tiny difference of 0.05% is observed, indicating excellent agreement between DNS and the κ-ε model. This result aligns with the findings of Hernández-López et al. [
9], who demonstrated that both approaches produce similar thermal profiles in turbulent regimes dominated by advective convection.
To examine this behavior in detail,
Figure 6 presents the dimensionless temperature profiles for high Rayleigh numbers in the range of 10
9 ≤ Ra ≤ 10
10. In these cases, the profiles obtained via direct numerical simulation (DNS) and those calculated using the κ-ε turbulence model are nearly identical, demonstrating strong agreement between the two methods under fully turbulent conditions. This behavior reinforces the previously observed trend, in which the discrepancy between the two models diminishes as the flow moves away from the transitional regime and advective convection becomes the dominant heat transfer mechanism.
These results are highly consistent with those reported by other authors in the literature [
7,
8,
18,
19,
33,
34], who concluded that under high Rayleigh number conditions and fully developed turbulent flow, turbulence models such as κ-ε can adequately capture the global thermal behavior without the need to resolve all flow scales, as DNS does. Similarly, Trias et al. [
33], applying DNS in a cavity with a geometric aspect ratio of 4 for Rayleigh numbers ranging from 6.4 × 10
8 to 1 × 10
10, observed that the temperature distribution exhibits clear stratification in the center and intense thermal gradients within the boundary layers—both of which are accurately described by DNS and RANS models, given that the mesh is sufficiently refined. Barakos et al. [
3] also emphasized that temperature profiles tend to stabilize in strongly convective regimes, and the differences between turbulence models and DNS decrease significantly if mesh refinement near the walls is adequate to resolve the thermal boundary layers. Overall, the excellent agreement observed in this study for Ra ≥ 10
9 validates the capability of direct numerical simulation to predict thermal distributions in differentially heated cavities and demonstrates that under fully turbulent conditions, both approaches converge to virtually equivalent solutions [
9,
13,
19].
This behavior also indicates that the global energy balance stabilizes once a sufficient resolution is achieved to capture the transport mechanisms in regions with the highest gradients, thereby validating the implemented numerical scheme.
Regarding the temperature contours,
Figure 7 shows those obtained from direct numerical simulation (DNS), while
Figure 8 displays the results generated using the κ-ε turbulence models. The analysis focuses on thermal behavior at high Rayleigh numbers in the range of 10
8 ≤ Ra ≤ 10
10, highlighting the evolution of thermal stratification, boundary layers, and the transition to the turbulent regime. Temperature contours for Ra values from 10
3 to 10
8 have been extensively studied in the literature (authors) and reported in a previous study by Molina-Herrera et al. [
2]. Within this Rayleigh range, the flow gradually transitions from conduction-dominated behavior to increasingly intense convection as Ra increases. It is well documented that the rise in the Rayleigh number makes thermal stratification and boundary layer thinning more pronounced. Since these earlier results have been validated in multiple studies, the present work focuses on the thermal behavior for Ra = 10
8 to 10
10, where the flow reaches a fully turbulent regime and heat transfer is predominantly advective.
These figures show that for Ra = 10
9, the temperature contours obtained with DNS and the κ-ε turbulence model exhibit significant thermal stratification, displaying intensified temperature gradients along the hot and cold walls. However, the κ-ε model slightly underestimates thermal dissipation near the walls, which aligns with the findings of Hernández-López et al. [
9], who noted that RANS models tend to reduce diffusive effects in the boundary layer region. Nevertheless, the similarity in the overall shape of the thermal contours indicates that both models effectively capture the impact of advective convection. Additionally, the studies by Trias et al. [
33] demonstrate that, from this Rayleigh number onward, the flow develops small instabilities within the thermal boundary layers, which DNS captures more accurately. In contrast, the κ-ε model produces smoother isotherms, suggesting that turbulent transport is slightly overestimated compared to the DNS solution. As the Rayleigh number rises to Ra = 10
10, thermal stratification becomes even more pronounced, with nearly horizontal isotherms in the center of the cavity. Both models agree that convection is the primary heat transfer mechanism, consistent with findings reported by various authors [
13].
A notable difference is that DNS captures the intensity of thermal gradients in the boundary layer more accurately. In contrast, the κ-ε model tends to smooth the profile in this region. This observation aligns with previous studies indicating that turbulence models often overestimate turbulent dissipation, resulting in an artificially thickened thermal boundary layer. Nevertheless, both models exhibit well-defined thermal recirculation zones in the upper region of the cavity, suggesting that the global flow structure is consistently represented in both simulations.
Trias et al. [
33] also reported that thermal stratification is nearly complete in this range of Ra, a characteristic evident in both simulations. This indicates that the difference between DNS and the κ-ε model in this regime is primarily quantitative, as the overall flow structure and thermal distribution remain comparable. For Ra = 10
10, the flow achieves a fully turbulent regime, with thermal stratification nearly complete at the center of the cavity [
13,
19]. DNS and the κ-ε model produce virtually identical isotherms in this scenario, confirming that advective convection entirely dominates heat transfer. Furthermore, the thermal gradients at the hot and cold walls are steep, and the thermal boundary layer is exceedingly thin. Fluid recirculation in the upper part of the cavity intensifies, indicating that thermal energy is rapidly transported upward, consistent with previous findings [
9,
13,
19]. The difference between the two models is minimal, with a discrepancy of less than 0.001% in the dimensionless temperature values at the hot wall. This implies that under this fully turbulent regime, the κ-ε model and DNS practically converge to the same solution. In terms of overall comparison, for Rayleigh numbers of 10
9 and 10
10, the κ-ε model tends to overestimate thermal dissipation in the boundary layer compared to DNS. For Ra = 10
10, the thermal structure is similar in both models, although DNS captures the boundary layer gradients more accurately.
Figure 9 illustrates the velocity profiles of the
Uy component along the horizontal
X-axis for Rayleigh numbers ranging from 10
3 to 10
10, obtained through direct numerical simulation (DNS) and the κ-ε turbulence model. The solid lines represent the results from DNS, while the dashed lines correspond to the predictions from the κ-ε model. For both models, it is noted that the magnitude of
Uy increases as the Rayleigh number rises, indicating a strengthening of convective flow within the cavity. For Ra ≤ 10
5, the velocity profiles obtained from DNS exhibit a parabolic shape near the walls, with peak values in those regions and a minimum at the center of the cavity, suggesting strong recirculation. As the Rayleigh number increases, starting from Ra = 10
6, the profiles gradually thin the boundary layer and transition towards nearly linear shapes, characteristic of regimes where convection strongly dominates over thermal diffusion. For Ra ≥ 10
8, the profile becomes almost linear, indicating a highly stratified and advective flow [
18].
In contrast, the velocity profiles predicted by the κ-ε model show significant differences. For Ra = 10
3, the profile is nearly linear. It fails to reproduce the expected parabolic shape, indicating a poor representation of laminar flow due to the excessive turbulent dissipation characteristic of this model. As Ra increases, the κ-ε model starts to capture parabolic-type profiles near the differentially heated walls, particularly for Ra ≥ 10
5. However, this behavior does not resemble the pattern observed with DNS; unlike the latter, the κ-ε velocity profiles maintain parabolic shapes even at high Rayleigh numbers, which contradicts the tendency toward linear profiles characteristic of the dominant advective regime. This discrepancy suggests that the κ-ε model struggles to capture the flattening of the central profile adequately and tends to produce pronounced velocity gradients near the heated walls, limiting its accuracy in predicting flow in differentially heated cavities. Overall, these results confirm that while the κ-ε model is helpful in highly turbulent regimes, its ability to represent flow dynamics accurately is limited in laminar and transitional regimes, where DNS provides a more precise description in both magnitude and shape of the velocity profiles [
9,
13,
19,
20].
Extending the analysis,
Figure 10 shows the velocity profiles of the horizontal component Ux along the cavity height (
Y), as predicted by direct numerical simulation (DNS) and the κ-ε turbulence model. As in the previous figure, solid lines represent DNS results, while dashed lines correspond to κ-ε predictions. The DNS-calculated profiles exhibit a symmetric, parabolic shape with well-defined peaks near the upper and lower walls (regions close to
Y = 0 and
Y = 1). This behavior is typical of laminar flows with established natural convection, generally observed for Ra ≤ 10
5, where convection begins to dominate over thermal diffusion. In this regime, intense flow is generated along the horizontal walls, with maximum velocities linked to the upward motion of hot fluid near the right wall and the downward motion of cold fluid near the left wall. As the Rayleigh number increases (Ra ≥ 10
6), a progressive reduction in the
Ux velocity magnitude occurs in the central region of the cavity. This trend continues with increasing Ra, and for very high values (Ra ≥ 10
9), the profiles tend to flatten, and the
Ux component approaches zero throughout the cavity height [
18]. This phenomenon can be explained by the intensification of the advective nature of the flow: as the Rayleigh number increases, motion within the cavity becomes increasingly concentrated in narrow thermal boundary layers adjacent to the vertical walls, significantly reducing the magnitude of horizontal flow in the cavity core. In this strongly stratified regime, natural convection generates dominant vertical flows, while horizontal components diminish, resulting in near-zero
Ux velocities outside the boundary layers [
18,
22,
23,
24].
In contrast, the velocity profiles predicted by the κ-ε model show notable discrepancies when compared to those obtained with DNS. For Ra = 10
3, the profiles are significantly smoother and do not capture the characteristic parabolic structures observed with DNS. This reduction in
Ux velocity component values results from the nature of the κ-ε model, which introduces turbulent kinetic energy dissipation (ε), particularly significant in laminar and transitional regimes. In these regimes, the production of turbulent kinetic energy (κ) is nearly zero due to the lack of strong fluctuations in the flow, causing dissipation to far exceed production and thereby reducing the intensity of the velocity profiles. As the Rayleigh number increases, the
Ux velocity values tend to become nearly linear, approaching values close to zero over much of the domain. This behavior is attributed to the predominant dissipation of turbulent kinetic energy (ε) and the low production of energy (κ), which restricts the development of intense turbulent structures in areas distant from the walls, especially in the cavity center. Furthermore, the formation of thin boundary layers near the walls, where convective flow prevails, concentrates velocity gradients in these regions. In contrast, the flow in the interior of the domain becomes more uniform [
13,
14,
19].
Furthermore,
Table 8 presents the maximum velocities of the dimensionless components
Ux and
Uy, calculated using direct numerical simulation (DNS) and compared with the values reported by De Vahl Davis [
1] for Rayleigh numbers Ra ≤ 10
6, as well as the results obtained by Goloviznin et al. [
8] for Ra ≥ 10
7. The maximum velocity values for each component were determined along lines perpendicular to the respective directions (i.e., Ux along a vertical line at X = 0.5 and Uy along a horizontal line at Y = 0.5), locating the local maxima within the domain, with their coordinates given in parentheses. Additionally, the table presents the average Nusselt numbers and their percentage deviations from the benchmark values of De Vahl Davis [
1] and Goloviznin et al. [
8].
Table 8 shows that direct numerical simulation, when implemented with mesh refinement along the differentially heated walls, provides results consistent with those of De Vahl Davis up to Ra = 10
6, with deviations below 1% in the Nusselt number, thereby validating the accuracy of the numerical model in the laminar regime. However, for higher Rayleigh numbers (Ra ≥ 10
9), the maximum velocity values tend to shift toward the region near the upper wall and closer to the hot wall T
h. This difference is attributed to the mesh refinement in the near-wall regions, which enables better resolution of thermal and velocity boundary layers, causing the zones of highest convective transport and maximum velocity to concentrate in narrower regions closer to the hot boundaries. Overall, the results confirm that the numerical strategy employed is suitable for both the laminar regime and the transition to turbulent flow, accurately capturing the maximum velocities and the characteristic flow structures of natural convection in differentially heated cavities.
Meanwhile,
Table 9 compares the dimensionless maximum velocities of the
Ux and
Uy components, obtained using direct numerical simulation (DNS) and the standard κ-ε turbulence model, for different Rayleigh numbers. The maximum values were computed in the same way as in the previous analysis, specifically along lines perpendicular to each velocity component. The results indicate that, for low Rayleigh numbers (Ra ≤ 10
5), the maximum velocities predicted by the κ-ε model exhibit significant discrepancies when compared to those obtained using DNS. This deviation can be attributed to the excessive dissipation of turbulent kinetic energy in regimes where turbulence effects are not yet fully developed, which limits the model’s ability to predict the flow pattern accurately. As Ra increases (Ra ≥ 10
7), the differences between both methodologies progressively decrease, suggesting that the κ-ε model becomes more accurate in capturing the main flow features once the turbulent regime is fully established. This behavior confirms that the κ-ε model tends to underestimate maximum velocities in laminar and transitional regimes, due to its formulation based on assumptions of fully developed turbulence. However, in fully turbulent regimes, its predictions closely match those of DNS in both magnitude and spatial location.
Finally,
Figure 11 presents the streamlines obtained for Rayleigh numbers of Ra = 10
9 and Ra = 10
10, calculated using direct numerical simulation (DNS) and the κ-ε turbulence model. For Ra = 10
9, the streamlines obtained with DNS reveal a fully turbulent flow characterized by the formation of well-defined vortices in the upper left and lower right corners of the cavity. These recirculation zones indicate intense convective activity associated with the interaction between thin thermal boundary layers and strong velocity gradients. The remainder of the cavity displays parallel and flat streamlines, reflecting a highly convective and stratified regime. For Ra = 10
10, a similar but more intense pattern is observed: the corner vortices become more compact, and the streamlines in the core of the cavity remain horizontal and densely clustered. This behavior is typical of flows at high Rayleigh numbers, where viscous dissipation is localized within thin boundary layers, while the fluid core moves almost inertially.
In contrast, the streamlines generated by the κ-ε model for the identical Rayleigh numbers do not demonstrate clear vortex formation in the cavity corners. Instead, a smoother and more symmetric configuration is observed, with streamlines extending almost continuously in the horizontal direction. This absence of localized vorticity can be attributed to the inadequate representation of small-scale velocity fluctuations, as the κ-ε model tends to overestimate turbulent kinetic energy dissipation (ε) and underestimate its production (κ). Consequently, the model excessively smooths the flow structures, hindering its ability to capture the recirculation mechanisms under high natural convection conditions accurately. This limitation impacts the model’s accuracy in predicting flow patterns in cavities where thin thermal boundary layers and dynamic stratification are significant factors.
It is also important to mention that in previous studies reported by Molina-Herrera et al. [
2], streamlines were analyzed for Rayleigh numbers ranging from 10
3 to 10
8, using both direct numerical simulation (DNS) and turbulence models. In these regimes, particularly for Ra ≤ 10
5, the streamlines obtained through DNS displayed a single dominant recirculation cell centered in the cavity, with smooth and symmetric trajectories indicating a well-organized laminar flow, primarily driven by conduction with the onset of natural convection. As the Rayleigh number increases to the range of 10
6–10
8, more complex structures begin to form in the corners, particularly in the lower right and upper left zones, reflecting the emergence of convective instabilities and the transition to more convective flows.
In this context, although it is acknowledged that buoyancy-driven flows at high Rayleigh numbers often develop complex three-dimensional structures, numerous studies have shown that two-dimensional (2D) simulations remain effective for accurately predicting overall quantities such as the Nusselt number and main circulation patterns. Wang et al. [
19] and Trias et al. [
33] have demonstrated that, with refined meshes and suitable numerical methods, 2D models can reliably reproduce global thermal behavior up to Ra ≈ 10
9. These simulations are especially valuable for comparing different numerical approaches (such as DNS and κ-ε) in studies where overall accuracy is prioritized over detailed representation of fine three-dimensional structures [
27,
28]. Specifically, the results presented here indicate that 2D DNS can effectively capture the formation and evolution of corner vortices, which the κ-ε model does not reproduce. This type of 2D analysis provides a solid foundation for assessing the relative performance of various turbulence models and for understanding global transport mechanisms, provided the conclusions remain within the limited scope of this dimensionality.