3.1. Average Flow and Heat Transfer Characteristics
To investigate the average flow and heat transfer characteristics on the blade surface, this section analyzes the distributions of the
,
, and Nusselt number (
) under different
. All the averaged results presented in this section, including the surface distributions and boundary layer profiles, were obtained by averaging the instantaneous 3D DNS data in both time and the spanwise (
z) directions. The spanwise averaging is appropriate since we employed periodic boundary conditions in the
z-direction, making these quantities statistically two-dimensional. For time averaging, the flow was first allowed to reach a statistically steady state at approximately
(where
is the flow-through time). Subsequently, time averaging was performed over an additional period of nine flow-through times to ensure adequate statistical convergence [
26].
Figure 5 shows the distributions of
and
on the pressure surface under various
conditions. As shown in
Figure 5a, the boundary layer initially experiences an adverse pressure gradient followed by a favorable pressure gradient. This general trend in the
distribution remains consistent over the range of
values. It is worth noting that in the interval
, the three higher
cases (
,
, and
) exhibit a sharp increase in
, which is primarily due to the pressure recovery caused by flow reattachment.
Combining the friction coefficient
curves shown in
Figure 5b with the data listed in
Table 3, it is evident that the flow separation positions vary with
. Specifically, the two lower
cases (
and
) do not exhibit flow separation. For the three higher
cases, as
increases, the separation point on the pressure surface boundary layer moves upstream. According to
Table 3, the separation points are located at approximately
, 0.41, and 0.40 for
,
, and
, respectively, whereas the corresponding reattachment points shift forward to approximately
, 0.66, and 0.65. This pressure recovery occurs because the reattachment of separated flow changes the streamline curvature near the wall. As a result, the near-wall fluid experiences reduced deceleration and may even regain some kinetic energy, leading to an increase in static pressure, and consequently, an increase in the pressure coefficient
. These reattachment positions roughly correspond to the regions where
shows a sharp increase, as previously reported [
39]. Under these higher
conditions, the lengths of the laminar separation bubbles on the pressure surface are similar, occupying roughly 25%, 26%, and 27% of the chord length
, respectively.
Figure 6 depicts the distributions of
and
on the suction surface for various
values. Compared to the pressure surface, the flow state on the suction surface is more complex. As shown in
Figure 6a, the
on the suction surface first decreases rapidly, after which the flow enters a pronounced adverse pressure gradient region. Under the influence of this strong adverse pressure gradient and curvature effects, the boundary layer separates. The separation positions under different
conditions show little variation; according to
Table 3, the separation points are located between approximately
and
. After boundary layer separation, the near-wall flow decelerates such that the wall pressure remains nearly constant, forming a “pressure plateau” region [
19] similar to that observed in two-dimensional flow (see
Figure 4). In
Figure 6b, the
values for the two lower
cases (
and
) remain negative, indicating no flow reattachment. In contrast, reattachment occurs for all three higher
cases. Consistent with the pressure surface observations, as
increases, the reattachment position on the suction surface also moves upstream [
9,
31].
Figure 7 shows the distribution of the
on the blade surfaces. As depicted in
Figure 7a, on the pressure surface, the
distribution exhibits similarities to the
distribution. Although flow separation occurs on the pressure surface under the three higher Reynolds number conditions (
,
, and
), the separation intensity is relatively weak. This weak separation has a limited impact on disrupting the qualitative similarity between the
and
distributions, with both still maintaining a similar streamwise evolution trend, which aligns qualitatively with the classical Reynolds Analogy [
40]. In contrast, the strong flow separation on the suction surface (
Figure 7b) significantly affects the heat transfer characteristics. As shown in
Figure 7b,
initially decreases as the boundary layer develops. When the flow enters the separation region,
drops markedly to its minimum value; however, once the separated shear layer undergoes transition,
rises sharply [
13]. For the three higher
cases (
,
, and
) in which reattachment occurs, the
values remain relatively high in the downstream reattachment region. For the two lower
cases (
and
), where reattachment does not occur, the
values in the latter half of the suction surface stay low as the flow remains separated.
To more intuitively demonstrate the effects of
on the flow and temperature fields,
Figure 8 and
Figure 9 show the ensemble-averaged velocity magnitude (
) contours and temperature (
) contours, respectively. All velocity values in
Figure 8 have been nondimensionalized using the inlet velocity
U. The cases with the lowest
(
) and the highest
(
) are selected as representative examples.
Comparing
Figure 8a,b, the significant influence of
on the size of separation bubbles on the suction surface is evident. At the lower
, a large-scale separation region covers an extensive area on the suction surface, with low-speed fluid occupying a large portion of the near-wall region in the mid and rear parts of the airfoil. At the higher
, the separation region is notably reduced, and the boundary layer on the pressure surface also appears thinner. These observations agree with the quantitative data in
Table 3 and the separation and reattachment behavior revealed by the
distributions in
Figure 5b and
Figure 6b.
The corresponding temperature fields in
Figure 9 also reflect these differences in flow structure. At the lower
(
Figure 9a), associated with the large separation bubble, a very thick thermal boundary layer forms on the suction surface, directly leading to the low
values observed in
Figure 7b. At the higher
(
Figure 9b), with the reduction in the separation bubble and the thinning of the velocity boundary layer, the thermal boundary layer thickness decreases significantly. This reduction results in increased near-wall temperature gradients and, consequently, higher
values (see
Figure 7b).
3.2. Velocity and Thermal Boundary Layers Analysis
The flow separation and heat transfer phenomena on the blade surface are directly influenced by the state of the near-wall boundary layer. This section provides a detailed analysis of the velocity and thermal boundary layers near the blade. First, it is necessary to precisely determine the extent of the boundary layer. For a traditional zero-pressure-gradient boundary layer on a flat plate, the nominal thickness
is defined as the normal distance from the wall where the fluid velocity
u reaches 99% of the free-stream velocity
U, i.e.,
[
41]. However, in complex wall flows with pressure gradients (such as on blade surfaces), this definition is often no longer applicable.
Taking the case with
as an example,
Figure 10 shows the ensemble-averaged velocity field at this
and marks six characteristic locations (L1 to L6) for boundary layer analysis. Among these, L1 (
), L2 (
), and L3 (
) are on the pressure surface, corresponding to laminar, separated, and reattached regions, respectively. Similarly, L4 (
), L5 (
), and L6 (
) are on the suction surface, corresponding to the laminar, separated, and reattached regions.
Figure 11 and
Figure 12 then present the ensemble-averaged tangential velocity (
) and ensemble-averaged temperature (
) profiles at these six characteristic locations, where the velocity is nondimensionalized by the inlet velocity
U, and the wall-normal distance
is nondimensionalized by
.
As shown in
Figure 11, near the edge of the boundary layer, the velocity profiles indicate that the
no longer asymptotically approaches a constant value. Unlike the flat-plate boundary layer where velocity approaches a uniform free-stream value, the edge velocity varies along the blade surface due to the pressure gradient and flow acceleration in the cascade passage. In contrast, as demonstrated in
Figure 12, the
in the profiles does asymptotically approach a constant at the edge of the thermal boundary layer. Therefore, the thermal boundary layer thickness
can still be defined by the traditional method, that is, as the normal distance from the wall where
reaches 99% of the inlet reference temperature
.
Existing methods for defining the flow boundary layer based on average vorticity [
42] or shear stress thresholds [
43] typically involve complex iterations, numerical integration, or differential calculations. These techniques increase computational complexity and can introduce additional errors and instabilities. In this study, a new method is adopted to calculate the flow boundary layer thickness
, based on the reconstruction of local inviscid velocity profiles [
27].
For incompressible steady flow, the total pressure
is given by
where
P is the static pressure,
is the fluid density, and the square of the velocity magnitude,
, is determined by the mean tangential velocity
and the mean wall-normal velocity
in the two-dimensional local coordinate system. Since
reflects the total energy (or work capacity) of the fluid, in regions far from the wall, the diminishing viscous effects cause
to approach a constant value. Therefore, the maximum value of
in this region is used as the inviscid reference total pressure:
. By applying the Bernoulli equation in the wall-normal direction, the square of the corresponding inviscid flow velocity
can be reconstructed:
Based on this, the flow boundary layer thickness
is determined by the more general condition:
Figure 13 shows the curves of the actual velocity squared,
, and the reconstructed inviscid velocity squared,
, as functions of the wall-normal distance
at the six characteristic locations L1–L6. In the figure, both velocity-squared terms are normalized by
.
As shown in
Figure 13, in the outer region of the boundary layer, the reconstructed inviscid velocity squared
agrees well with the actual viscous velocity squared
. Thus, the flow boundary layer thickness
is clearly defined as the wall-normal distance at which
reaches 98% of
. The square markers in
Figure 13 indicate the calculated boundary layer thickness
at the six representative locations. The results show that on the pressure surface
, and on the suction surface
. This trend is consistent with the physical expectation that the flow boundary layer typically grows thicker in the streamwise direction. Therefore, this new method conveniently and consistently determines the flow boundary layer thickness in both attached and separated regions.
In velocity boundary layer theory, the displacement thickness
and momentum thickness
are two important integral measures that characterize the boundary layer from different perspectives [
39,
41]. The displacement thickness
measures the mass flow deficit due to the boundary layer velocity profile, while the momentum thickness
reflects the momentum loss. In this method, these quantities are defined as
,
, where
is the nominal boundary layer thickness obtained from the local inviscid velocity profile reconstruction,
is the mean tangential velocity at the boundary layer edge, and
is the wall-normal distance. Combining these, a dimensionless shape factor
H is defined:
, which serves as an important indicator for assessing the boundary layer state and predicting transition. The velocity boundary layer parameters and thermal boundary layer thickness on the pressure surface for different
are shown in
Figure 14, where the
,
, and
are all nondimensionalized using the axial chord length
.
Figure 14a indicates that the displacement thickness
at the two lower
is smaller than at the three higher
, which is attributed to flow separation in the latter, causing greater mass loss. As the
increases, the peak value of
decreases, suggesting a weakening of the separation. In
Figure 14b, the momentum thickness
increases slowly along the flow direction at the lower
. However, for the three higher
,
exhibits a sharp increase near
, indicating a rapid rise in near-wall momentum loss typically associated with transition phenomena [
44].
Figure 14c shows that the initial value of the shape factor
H is close to 2.5, a typical value for an attached laminar boundary layer [
41,
45]. At the two lower
,
H remains nearly constant in the first half of the flow and begins to decrease after
, yet remains relatively high, indicating an essentially laminar state on the pressure surface. For the three higher
,
H initially changes slowly and then exhibits a pronounced peak. When the separated shear layer becomes unstable and triggers transition, intensified momentum exchange causes a rapid increase in
, which in turn leads to a sharp decrease in
H. Consequently, the peak in
H can be regarded as the approximate location of transition onset [
46]. Based on these peaks, the transition locations for
,
, and
are approximately at
,
, and
, respectively, indicating that the transition position moves upstream with increasing
. In the region
, the
H values for these higher
tend to stabilize at levels lower than typical turbulent boundary layers (around 1.4–1.7) [
45], implying that the boundary layer has developed into a turbulent state in the rear part of the pressure surface. Correspondingly,
Figure 14d shows that after transition at higher
, the thermal boundary layer thickness
is significantly greater than the laminar thermal boundary layer thickness observed at the two lower
due to enhanced turbulent mixing. For the suction surface,
Figure 15 presents the boundary layer parameters and thermal boundary layer thickness at different
.
Figure 15a shows that due to significant separation, the displacement thickness
is markedly increased, reflecting substantial mass flow loss. However, as
increases, the degree of separation weakens and
correspondingly decreases.
Figure 15b shows that, similar to the pressure surface, even with separation, the momentum thickness
increases slowly before transition because the initial momentum loss is relatively small; once transition occurs, momentum loss intensifies and
increases rapidly. As indicated in
Figure 15c, the initial value of
H on the suction surface is also approximately 2.5, typical of a laminar boundary layer. Since flow separation occurs on the suction surface for all investigated
, the instability of the separated shear layer ultimately results in turbulent transition, and the
H curves for all cases exhibit distinct peaks. For the two lower
(
and
), the transition peaks occur at approximately
. For the three higher
(
,
, and
), the transition locations are at
, 0.75, and 0.74, respectively, indicating a slight upstream shift with increasing
. For
and
, after transition, the
H value stabilizes around 2. In contrast, for
and the two lower
, the reattachment point is further downstream (or even incomplete), so
H continues to decrease toward the trailing edge without stabilization, suggesting that the flow is still transitioning toward turbulence. This is attributed to the combined effects of the upstream large-scale separation history and the insufficient turbulent development length offered by the finite blade chord. Consequently, for these lower
cases, the boundary layer on the suction surface fails to achieve a fully developed turbulent state by the trailing edge of the blade. Regarding the thermal boundary layer (
Figure 15d),
increases slowly before separation and grows rapidly thereafter. Unlike on the pressure surface, on the suction surface, the thermal boundary layer thickness
decreases monotonically over the entire streamwise range as the
increases.
3.3. Turbulent Vortex Structure and Heat Transfer Enhancement Mechanism
Based on the preceding analysis of the average heat transfer characteristics and the boundary layer state on the blade surface, it is observed that once the flow transitions to a turbulent state, its convective heat transfer efficiency is significantly enhanced compared to that in the laminar region. To elucidate the physical mechanism underlying this phenomenon, an in-depth investigation of the three-dimensional flow structures and the associated turbulent characteristics within the turbulent boundary layer is essential. Because the flow and heat transfer characteristics exhibit a monotonic trend with increasing within the studied range, the most representative conditions were selected for presentation. For the pressure surface, conditions at , where transition has occurred and turbulent structures were formed, and the highest Reynolds number are presented. For the suction surface, the lowest Reynolds number , which exhibits large-scale separation, and the highest Reynolds number , which shows reattached flow, are presented to fully demonstrate the range of influence of the effects.
To effectively identify three-dimensional vortex structures, this study employs the Rortex method proposed by Liu et al. [
28]. This method decomposes the vorticity vector into a Rortex vector
, representing rigid rotation, and a non-rotating shear vector
. The Rortex vector is expressed as
where
,
, and
are the unit vectors in the Cartesian coordinate system
, respectively. The magnitude of the Rortex vector,
, which characterizes the strength of the rigid rotation, is defined as follows:
here,
and
are computed as
where
denote the velocity components in the principal coordinate system
. Compared with traditional vortex identification methods, the isosurfaces of
provide a more accurate characterization of vortex structures by eliminating shear effects and retaining only the pure rigid rotation of the Rortex as a clear vortex identifier [
47].
Figure 16 illustrates the instantaneous flow structure and thermal field characteristics on the pressure surface under two
conditions (
and
). The three-dimensional vortex structures are visualized by the
isosurface and are colored by the temperature
. To reveal the direct correlation between these structures and the near-wall thermal field, the figure simultaneously displays the instantaneous
distribution and velocity vector field on two characteristic cross-sections. The spanwise cross-section S1, located at
, was chosen to avoid the influence of periodic boundaries and to capture typical spanwise flow characteristics. The wall-parallel cross-section S2, located at
, corresponds to a
in the turbulent region, placing it within the buffer layer where near-wall streak structures are most prominent.
In
Figure 16a for the
case, early two-dimensional Kelvin–Helmholtz (K-H) vortices appear in the region
–0.7. Subsequently, these K-H vortices become unstable and evolve into three-dimensional structures [
19]. As the flow reattaches, hairpin vortex structures, similar to those observed during natural transition, form near the wall. These structures consist of streamwise streaks in the near-wall region and arched vortex heads further away from the wall [
48]. In this manner, K-H vortices and hairpin vortices are the primary vortex structures on the pressure surface. Section S1 displays the instantaneous morphology of the thermal boundary layer at the
position, along with the cross-sectional features observed when K-H or hairpin vortex heads pass through this plane (indicated by white dashed circles). The presence of these vortex structures significantly distorts the isotherms, resulting in localized downward displacement of hot fluid. Section S2 more clearly reveals the thermal effects of the ejection and sweep events induced by hairpin vortices in the near-wall region, that is, the upwash motion beneath the hairpin vortex legs ejects near-wall cold fluid, while sweeps on the outer sides of the legs transport mainstream hot fluid toward the wall. This is manifested as alternating blue (cold) and red (hot) streaks on the S2 plane.
A comparison of
Figure 16a,b indicates that at
, the onset of K-H vortices on the pressure surface shifts upstream. The flow visualization also reveals that the hairpin vortex structures exhibit smaller scales and denser spatial distribution. Correspondingly, the temperature fields on planes S1 and S2 exhibit smaller-scale, more intricate mixing patterns, signifying intensified turbulent mixing at the higher
. The persistent influence of these instantaneous vortex structures and their induced ejection and sweep events is clearly reflected in the turbulent statistics.
Figure 17 presents the ensemble-averaged turbulent statistics near the pressure surface for the two
, including the root mean square (RMS) of velocity fluctuations (
and
), RMS of temperature fluctuations (
), Reynolds shear stress (
), and turbulent heat fluxes (
and
). These statistical quantities are all presented in dimensionless form: the RMS of velocity fluctuations are normalized by
U; Reynolds stress is normalized by
; and turbulent heat fluxes, which are the product of dimensionless velocity fluctuations and dimensionless temperature fluctuations, are inherently dimensionless.
In
Figure 17a, for the
case, the high-value regions of the RMS velocity and temperature fluctuations (
,
, and
) are primarily concentrated in the region following turbulent transition, with peaks located near the wall. This corresponds to the active region of hairpin vortices shown in
Figure 16a, reflecting the strong fluctuations induced by these vortex structures. In the region where
, the Reynolds shear stress
exhibits negative values in the near-wall region, while it approaches zero farther away from the wall. This distribution indicates the transport of turbulent momentum toward the wall, which is a typical characteristic of an attached turbulent boundary layer closely associated with bursting and sweeping events [
45]. Similarly, the normal turbulent heat flux
also shows negative values in the near-wall region and exhibits a distribution pattern similar to that of
. In contrast, the streamwise turbulent heat flux
is positive, indicating a tendency for streamwise velocity fluctuations and temperature fluctuations to be positively correlated, with its peak occurring closer to the wall than those of
and
. When the
increases to
(
Figure 17b), the peak locations of all statistical quantities shift upstream, in line with the earlier transition location. Simultaneously, the peak intensities are significantly enhanced, indicating that both turbulence intensity and turbulent transport efficiency (for momentum and heat) increase with
.
The instantaneous flow structures and associated thermal field characteristics on the suction surface under two
conditions, namely
and
, are depicted in
Figure 18. The three-dimensional vortex structures are visualized by the
isosurface and colored by the temperature
. The figure includes four sections for the present analysis: S3, S4, S5, and S6. Section S3 is a spanwise plane located at
; this position was chosen over the
used for the pressure surface as it better captures the characteristic vortex structures on the suction side. Section S4 is a near-wall parallel plane at
. Finally, considering the large-scale flow separation, two streamwise-normal sections, S5 (at
) and S6 (at
), are also included.
In
Figure 18a, at the lowest
(
), large-scale separation produces prominent spanwise K-H roll-up structures originating from the instability of both the mainstream and the separated shear layer. Their spanwise cross-sectional features are displayed in section S3 (the leftmost white dashed circle indicates the K-H vortex core region). Their rotation drives the early transport of high-temperature fluid toward the wall and the lifting of low-temperature fluid, in agreement with the findings on the pressure surface. As the flow develops downstream, the K-H vortices break down into a series of complex three-dimensional vortex structures, primarily consisting of quasi-streamwise vortices and distorted arch-like vortices. The near-wall flow is dominated by quasi-streamwise vortices, which manifest as temperature streaks in the wall-parallel section S4 (i.e., band-like distributions formed by the stretching of high- and low-temperature fluids along the streamwise direction). In contrast to the pressure surface, the suction surface separation zone exhibits smaller-scale and fragmented vortex structures. These fragmented vortices significantly shorten the streamwise streaks. Additionally, their random orientations and three-dimensional interactions generate spanwise streaks. In the streamwise sections S5 and S6, the red high-temperature regions become more prominent, reflecting further enhancement of thermal mixing and thickening of the thermal boundary layer. These sections simultaneously display the vortex core structures and the associated complex temperature distributions (as indicated by the white dashed circles), where strong rotational motion significantly affects the
field. Vortex rotation typically occurs in regions where the instantaneous
isolines change abruptly, and these rotational zones are often accompanied by interfaces between high- and low-temperature fluids, indicating that the rotational effect locally increases the instantaneous temperature gradient
.
When the
increases to
(
Figure 18b), the flow on the suction surface reattaches, and the turbulent mixing layer becomes more confined to the near-wall region. Consequently, in sections S3, S5, and S6 the overall thickness of the mixing layer is reduced compared with the low
case. Despite this reduction, turbulent activity within the layer is significantly enhanced due to the increased
. The vortex structures exhibit greater fragmentation and complexity, with the scale of K-H vortices decreasing and small-scale, distorted arch-like vortex structures becoming more prevalent. Moreover, the temperature field in the near-wall section S4 displays finer, more chaotic streak patterns. Notably, in sections S5 and S6, although the mixing layer is thinner, the proportion of red high-temperature regions increases, particularly near the vortex cores (as indicated by white dashed circles). This suggests that under high
and reattachment conditions, the vortex rotation and turbulent mixing in the near-wall region are significantly enhanced. These enhanced turbulent activities entrained and transported high-temperature mainstream fluid toward the wall more effectively. Compared to the low
case with large-scale separation, the thermal boundary layer at high
is inherently thinner. Within this thinner boundary layer, the intensified turbulent transport results in steeper local temperature gradients and more intense heat exchange near the wall.
The ensemble-averaged turbulent statistics on the suction surface for
and
are shown in
Figure 19.
In
Figure 19a, at the lowest
,
, the large-scale open separation produces a significantly elevated shear layer, with the peak regions of the RMS velocity fluctuations (
and
) located far from the wall (concentrated at
–0.08). This corresponds precisely to the core region of vortex breakdown and interaction observed in
Figure 18a. The strong fluctuations away from the wall are a direct manifestation of the separated shear layer instability. However, the RMS temperature fluctuation (
) exhibits a distinctive dual-peak distribution: one peak within the elevated shear layer (
–0.08) associated with large-scale mixing and K-H instability, and another stronger peak in the downstream near-wall region (
). The formation mechanism of the latter differs from that of the former; as the separated shear layer develops downstream and undergoes transition, the resulting small-scale turbulent structures can penetrate into the near-wall region. Because of the extremely steep mean temperature gradient near the wall, even small wall-normal velocity fluctuations can generate intense temperature fluctuations, contrasting with velocity fluctuations which are suppressed by the no-slip condition. Notably, the distribution of the Reynolds shear stress
, although predominantly negative, exhibits a weak positive peak along the shear layer (roughly corresponding to the location of the K-H vortices in
Figure 18a). In attached boundary layers, such as on the pressure surface,
is typically negative, indicating momentum transport toward the wall. Here, the positive region indicates momentum transport in the opposite direction within the separated shear layer, driven by the entrainment and flapping motions of large-scale K-H vortices—that is, momentum transport from the near-vortex-core region toward the outer edge of the shear layer. Additionally, the normal turbulent heat flux
exhibits a similar dual-peak pattern, with peaks observed in both the shear layer and the near-wall region, reflecting the combined effects of
fluctuations and the temperature gradient distribution discussed above.
In contrast, at the higher Reynolds number
(
Figure 19b), the peak intensities of all turbulent statistical quantities are markedly enhanced, and their peak regions shift closer to the wall. The positive peak region of
is smaller, and the negative peak is also located nearer to the wall, reflecting intense near-wall momentum exchange driven by smaller-scale vortex structures. This enhanced momentum exchange enables the flow to overcome the adverse pressure gradient and reattach. Furthermore, the peaks of the turbulent heat fluxes
and
are stronger and occur closer to the wall, indicating significantly improved heat transport in the near-wall region compared to the low
case, in agreement with [
49].
In summary, the vortex structures on the blade surface (such as K-H vortices, quasi-streamwise vortices, hairpin vortices, and arch-like vortices) play a crucial role in promoting mixing between the mainstream and the near-wall fluid by inducing ejection and sweep events. The cumulative effect of these instantaneous flow behaviors is ultimately reflected in the turbulent statistics, which show enhanced velocity and temperature fluctuations, increased Reynolds stresses, and higher turbulent heat fluxes. As the
increases, these vortex structures become smaller in scale and more densely distributed, leading to enhanced peak values of turbulent statistics closer to the wall and thereby effectively improving the convective heat transfer efficiency on the blade surface. These detailed physical mechanisms revealed under zero-FST conditions establish an essential baseline for understanding the flow and heat transfer phenomena in blade passages. To extend these insights to practical compressor applications, it is important to acknowledge that FST is ubiquitous in real environments and can modify the observed flow characteristics. Specifically, FST promotes earlier transition through bypass mechanisms, injects energy into separated shear layers to accelerate reattachment and reduce separation bubble size, and causes vortical structures to become more fragmented and complex, thereby altering the friction and heat transfer distributions on blade surfaces [
19]. Future studies will systematically investigate the coupled effects of
and FST to predict cascade flow and heat transfer under realistic operating conditions more accurately.