Abstract

High-pressure turbine blade (HPTB) shrouds suffer manufacturing variations in both platform alignment and inter-platform gap width. Compared to hub endwalls, the aerothermal effects of shroud platform steps and gaps has had little attention, which introduces uncertainty in the sentencing of such manufacturing variations. This article presents a shroud step sentencing correlation developed using a parametric quasi-2D (Q2D) model of a shroud endwall step. The use of a Q2D model follows from the study of a 3D steady Reynolds-averaged Navier–Stokes (RANS) simulation matrix of engine-representative platform steps and gap widths, based on a sample of scanned HPTB castings and finished parts. This study showed that the aftchord shroud step flow is Q2D and resembles canonical step flow with enhanced heat transfer at the reattachment point. The shroud step sentencing correlation is tested on the platform steps in the simulation matrix giving prediction errors below 20% for the majority of cases. Finally, the correlation is tuned using experimental data to mitigate the uncertainty associated with RANS simulations of separated flows.

1 Introduction

Shrouded high-pressure turbine blades (HPTBs), such as shown in Fig. 1(a), reduce tip leakage flow using a platform with fins and fences. Manufacturing variability leads to misalignments between adjacent shrouds and variable inter-platform gaps. Shrouds are exposed to some of the highest loads in the engine, yet unlike hub platforms, the aerothermal effect of shroud platform manufacturing variations is nearly absent from the literature. Uncertainty in the sentencing of shroud manufacturing variations risks both scrapping good parts and using parts prone to wear, leading to early maintenance, with reported costs of up to $2M per shop visit [1]. To help reduce this uncertainty, this article presents a correlation for the shroud step heat transfer, which can be used to triage parts within manufacturing line timescales. Figures 1(b) and 1(c) highlight the key finding that the aftchord shroud step flow is quasi-2D (Q2D) and resembles the corresponding canonical step flow with heat transfer enhancement at the reattachment point. It will be shown that prediction errors below 20% are obtained for the majority of cases when the correlation is tested on 3D HPTB shroud endwall steps.

Fig. 1
Shroud platform (indicated on a full blade) with corresponding Nu contours, demonstrating the Q2D flow over a backward-facing step in the aftchord region
Fig. 1
Shroud platform (indicated on a full blade) with corresponding Nu contours, demonstrating the Q2D flow over a backward-facing step in the aftchord region
Close modal

1.1 Literature Review.

Prior studies on manufacturing variations can be divided into in-depth studies of a specific manufacturing variation applied to stylized geometry such as a cascade [24] and Monte Carlo simulations of the overall manufacturing variability occurring on the part of interest, represented by either a parametrized model [57] or principle component analysis modes [8,9]. In this first category, Cardwell et al. [4] studied platform misalignments in a nozzle guide vane cascade with rim seal cavity flow and no inter-platform purge flow. They showed that the cooling effectiveness decreases in the aftchord region for a local forward-facing (FWD) step and increases for a backward-facing (BWD) step. This has also been found by Hada and Thole [10] using Reynolds-averaged Navier–Stokes (RANS) simulations. Lange et al. [11] studied the influence of hub platform steps including both rim seal cavity flow and inter-platform purge flow using IR measurements and RANS simulations. They observed a decrease of area-averaged Nu on the aftchord suction side (SS) endwall of 4% for a local backward-facing step and an increase of 2% for a forward-facing step. The largest change in local Nu was observed at the throat, where the inter-platform flow gets ejected on the SS platform. The same observations where made by Zhang et al. [12] using RANS simulations.

The aligned shroud endwall flow study by Lehmann et al. [13] on an HPTB with an idealized three fin shroud showed that, as opposed to hub endwalls, main passage fluid gets ingested in the inter-platform gap over the entire chord due to the decrease in static pressure caused by the fins. Rushton [14] and Wallis et al. [15] showed that also the secondary flows differ from the hub endwall. The horseshoe vortex forms in the shroud inlet cavity and emerges onto the endwall at midpitch, away from the potential field of the leading edge (LE). The pressure side (PS) leg of the horseshoe vortex immediately gets swept toward the suction surface due to the cross-passage pressure gradient. Downstream of the swept vortex, Lehmann et al. [13] showed that the shroud endwall flow is initially aligned with the wedge face and from midchord onward traverses the platform from pressure to suction side, where the near-wall PS fluid gets ingested in the inter-platform gap. As a result, a new boundary layer forms on the aftchord SS endwall with a corresponding local maximum in Nu at the SS platform edge, as shown by IR measurements. Lehmann et al. [16] also studied the influence of gap width variations on shroud heat transfer, but the platform was not considered.

Even though the discussed prior art provides a good overview of the influence of hub platform steps on aerothermal performance, the different leakage and secondary flows on shroud endwalls suggests that there is an opportunity for studying shroud platform steps with a view to improved sentencing of these parts.

1.2 Outline.

In this article, the influence of engine-representative shroud platform manufacturing variations on aerothermal performance are investigated using steady RANS simulations of a three passage engine part HPTB model. A step sentencing correlation for the reattachment Nu is then developed using a Q2D parametric model of the shroud endwall. The performance of the correlation is demonstrated against the full 3D flow field. Finally, the limitations of the RANS simulations used to build the correlation are addressed using experimental data.

2 Methodology

First, the measurement of step heights and gap widths on two samples of 3D scanned blades is explained. Next, the numerical and experimental methods are discussed in turn.

2.1 Recreating Engine-Representative Manufacturing Variations.

Platform step heights and inter-platform gap widths are measured on structured light scans of 100 castings and 26 finished parts of an HPTB of a modern aircraft engine (the latter has been used in the study by Lee et al. [17]), scaled into the hot reference frame. The platform step height and inter-platform gap width are measured at the LE, 25%Cx, 50%Cx, and 75%Cx for all possible blade passages that can be created from the sample. Figures 2(a) and 2(b) show that the step height, H, is measured on the intersection of an axial cutting plane and the shroud platforms at each axial location, whereas the gap width, w, is measured on the intersection of a plane cutting both wedge faces. For clarity, only the axial plane at 25%Cx is shown in Fig. 2(a). Since the wedge faces of the platform are machined on the parts after the casting stage, the inter-platform gap width cannot be measured in the casting sample.

Fig. 2
Measurement and reconstruction of shroud platform manufacturing variations
Fig. 2
Measurement and reconstruction of shroud platform manufacturing variations
Close modal

Figure 2(c) shows the statistical distribution of the measured step heights, normalized by blade span, S. A positive value indicates that the SS platform is more radially outward than the PS platform, as is the case shown in Fig. 2(a). Therefore, a positive value represents a forward-facing step (shown in Fig. 6(d)) at the LE where the flow crosses the step from SS to PS, and a backward-facing step (shown in Fig. 6(b)) for the locations further downstream, where the flow crosses the platform from PS to SS. The statistical gap width distribution is shown in Fig. 2(d). The standard deviation is larger for gap widths, illustrating the larger manufacturing variability. Note that even though negative gap widths occur in the hypothetical pairs, the shrouds would simply touch in reality.

A test matrix of three endwall misalignments and three inter-platform gap widths is created to study the corresponding effect on aerothermal performance. The manufacturing variations are chosen based on the statistical distribution shown in Figs. 2(c) and 2(d), such that their probability of occurrence can be determined. First, the histograms are compared to the equivalent normal probability density function, calculated using the respective mean, μ and standard deviation, σ. As such for the finished parts step height histograms, a multimodal distribution was needed, which suggests that the step heights are created by two or three distinct mechanisms. For the casting scan step height and the gap width distributions, however, a good match is obtained with a Gaussian.

The step height used in the following study is a μ+4σ variation, indicated in Fig. 2(c), referred to as the 4σ backward-facing step. This represents a large step and somewhat rare occurrence, with a probability to occur once in 240 engines. The same deviation mirrored around an aligned endwall is studied as well, referred to as the 4σ forward-facing step. A test case with an aligned endwall provides a reference. The studied gap widths consist of the nominal and a μ±σ gap, indicated in Fig. 2(d), of which the latter have a probability to occur in every engine.

Figures 2(e) and 2(f) illustrate how the manufacturing variations are applied to the StereoLithography (STL) geometry of the design-intent blade in matlab. The perturbations are applied using control points, shown in red, located where the manufacturing variations have been measured. An additional control point is added at the trailing edge (TE) to smooth the profile of the desired perturbation.

2.2 Numerical Methods.

Steady RANS simulations of a three passage high-pressure turbine (HPT) rotor model, a Q2D shroud endwall step model, and the internal geometry of the step heat transfer rig discussed in Sec. 2.3 are conducted to calculate Nu. The simulations are run in star-ccm+ using the kω shear stress transport (SST) turbulence model, where Nu is calculated using the two-run method. Meshing is performed using the automated unstructured polyhedral mesher in star-ccm+. Since the heat transfer coefficient (HTC) is a local property, grid refinement is applied in the vicinity of the step. A grid independence study is performed on the Q2D shroud platform step model for the mesh controls listed in Table 1. The grid-independent settings, shown in bold in Table 1, are obtained at the penultimate refinement, indicated by N-1, from which further refinement, indicated by N, leads to a relative change in reattachment Nu smaller than 2%. Note that the near step mesh refinement is mainly controlled by the local prism layer. The domain creation and boundary conditions for each model are now discussed.

Table 1

Normalized grid-independent mesh controls

Refinement for ΔNu<2%N-1N
Surface mesh triangle target size0.05 H0.025 H
Number of prism layers3040
Prism layer total thicknessδ991.25δ99
Wall y+<1.0<0.5
Refinement for ΔNu<2%N-1N
Surface mesh triangle target size0.05 H0.025 H
Number of prism layers3040
Prism layer total thicknessδ991.25δ99
Wall y+<1.0<0.5

Note: Grid-independent settings shown in bold are obtained at the penultimate refinement.

2.2.1 Three Passage High-Pressure Turbine Rotor Model.

The domain is created using the design-intent geometry of the blades transformed into the hot reference frame, including all shroud geometry but discarding cooling features except the shroud sealing flow. The tip gap is set to a typical mid-power value. Figure 3 shows the domain creation in star-ccm+, where four blades are subtracted from an empty periodic flow domain matching three blade passages. Manufacturing variations are applied to the red zone on the shroud platforms of the central passage. Using the grid-independent settings, a 51 million cell mesh is created.

Fig. 3
Illustration of the domain creation
Fig. 3
Illustration of the domain creation
Close modal

The three passage HPT rotor model is run with temperature ratios of 0.92 and 0.96 using boundary conditions extracted from a full stage simulation at mid-power.

2.2.2 Q2D Shroud Step Model.

The parametric study is performed on a Q2D step model based on the streamwise cut in Fig. 1(b). The step height is equal to 0.75%S, which is approximately equal to the mean of the measured step heights at 50%Cx. The gap width equals half the step height and is comparable to the mean measured gap width of around 0.375%S. The domain extends for 20H both upstream and downstream of the step. Upstream, this distance is used to let the boundary layer develop and downstream to capture the reattachment point well within the domain. The Q2D step domain has a depth of 2.5H, such that flow angle variation with respect to the step can be simulated. The height of the domain is extended to 250H such that a step height to passage-width ratio representative of commonly occurring shroud platform steps is created.

The walls forming the step at the bottom of the domain are isothermal, with the temperature ratio of each run given in Table 2. A slip wall is applied on the top of the domain, and the sides are periodic to allow for flow angle variations. The bottom of the gap can be a wall or a mass flow outlet, depending on whether gap leakage flow is simulated. A wall-normal profile for both total pressure and turbulence properties is specified at the inlet in order to define the inlet boundary layer. Inlet boundary conditions are extracted from a streamwise slice through the aftchord shroud platform step in the three passage HPT model. A summary of the freestream flow conditions in the Q2D domain is provided in Table 2, where Tu represents the turbulence intensity in the streamwise direction and TuL represents the turbulence length scale.

Table 2

Flow conditions in the Q2D shroud step domain

Q2D shroud step flow conditions
ReyH7600
Mis0.69
Tw,i/T0,in0.92 and 0.96
Tuin(%)6.0
TuL,in(%H)100.0
Q2D shroud step flow conditions
ReyH7600
Mis0.69
Tw,i/T0,in0.92 and 0.96
Tuin(%)6.0
TuL,in(%H)100.0

2.2.3 Step Heat Transfer Rig.

The internal geometry of the 2D step heat transfer rig described in Sec. 2.3 is modelled from the turbulence grid to the last pressure tapping in the test section. The experimental pressure and temperature data are used to set the boundary conditions, and the freestream turbulence at the step edge is matched to the experimental value. The step Reynolds number is 12,000, and the wall temperature ratios are 0.89 and 0.93, respectively.

2.3 Experimental Methods.

Experimental Nu data are collected using a 2D step heat transfer test rig, shown in Fig. 4. The rig is designed to conduct a transient heat transfer experiment, where the HTC is calculated by measuring the thermal response of a test plate to an increase in freestream temperature of 85 K by heat addition through two 15 kW heater meshes. The flow is guided to the test section using an 8 to 1 contraction. The boundary layer is bled on all four walls of the test section in order to reduce flow acceleration by boundary layer blockage. The onset of boundary layer transition on the test plate is fixed using a Braslow trip [18]. The freestream turbulence intensity, Tu, induced by a turbulence grid has been measured using a hotwire and is equal to 4.8% in the vicinity of the step. The flow field in the test section is fixed by choking the flow downstream of the test section.

Fig. 4
The 2D step heat transfer rig in the Whittle lab
Fig. 4
The 2D step heat transfer rig in the Whittle lab
Close modal

The temperature on the test plate is measured using a FLIR SC7000LW IR camera, which has a noise equivalent temperature of 25 mK and a measurement frequency of 200 Hz. The corresponding heat flux is calculated using the impulse-response method of Oldfield [19], which assumes 1D heat transfer in a semi-infinite substrate. Therefore, the step plates are designed such that errors introduced by violating these assumptions do not reach the reattachment point. This has been achieved by manufacturing the test plates out of polycarbonate, with measured thermal effusivity of 543Jm2Ks12. This allows to measure the thermal response of the test plate during 20 s, while limiting the lateral conduction errors to an area within 1.5H from the step edge.

The HTC is calculated using floating regression [20] of the measured temperature data, normalized by the time accurate Taw increase, and the corresponding heat flux data at each pixel. The experimental uncertainty is estimated to be 5% based on the Monte Carlo analysis of uncertainties performed by Playford [21].

The 2D HTC contours are spanwise-averaged between the boundaries indicated in Fig. 5, where Rex is measured from the plate LE, to avoid camera occlusion or corner flow effects.

Fig. 5
Experimental NuH contours on the BWD step
Fig. 5
Experimental NuH contours on the BWD step
Close modal

3 The Impact of Shroud Platform Steps on Aerothermal Performance

In this section, the results of the simulation matrix are discussed for each applied platform step, starting from the aligned endwall. The influence of changing the inter-platform gap width is discussed with respect to every considered platform step.

Since the developed Q2D correlation will be tested against the results of this 3D simulation matrix, ideally the same methods should be used. Although an experimental parametric study of the Q2D shroud step model has been conducted, recreating the 3D endwall platform step flow in an experiment is impractical. Since the focus lies on wall-bounded shroud step flow, the closest alternative numerical method to RANS is a wall-resolved large eddy simulation (LES), of which the computational cost is comparable to a direct numerical simulation (DNS) [22]. A recent DNS study on a cascade style single passage HPT stage at engine-representative Re required an 8.2 billion cell mesh [23], illustrating the prohibitive computational cost for use in this type of parameter study. Hence, the 3D simulation matrix and Q2D shroud model are run using steady RANS and the associated uncertainties are mitigated using experimental results, which are presented in Sec. 6. For the current discussion, it is sufficient to know that the experiments showed that a correction factor of 1.37 is required on the predicted reattachment Nu, where the flow resembles canonical forward-facing step flow. No correction factor is required for backward-facing step flow.

3.1 Aligned Endwall Flow Field.

The endwall flow in the vicinity of the wedge face depends on the pressure difference over the inter-platform gap, which is shown in Fig. 6(a), where Cp=(pinp)/12ρinVrel,in2 with the subscript, in, referring to the rotor inlet conditions. The pressure drop downstream of the first fin initially causes flow to be driven to the over-shroud region. As the Cp difference decreases between 20 to 40%Cx, the ingested main passage flow migrates parallel to the wedge face, as shown in the gap flow schematic in Fig. 6(a). At 40%Cx, the pressure difference reverses and the entrained gap flow gets driven back toward the main passage. From 50%Cx onward, the pressure drop downstream of the second fin leads to the ingestion of main passage flow.

Fig. 6
Cp difference over the inter-platform gap and streamlines in the aftchord region
Fig. 6
Cp difference over the inter-platform gap and streamlines in the aftchord region
Close modal

The surface streamlines shown in Figs. 7(a)7(c) confirm the conclusions drawn from the gap Cp distribution. In the LE region, the nearly wedge face aligned flow gets ingested into the inter-platform gap from both platforms. At midchord, the ejected inter-platform gap flow reattaches on the SS endwall, indicated by reattachment line A. The reattachment occurs further downstream than 40%Cx where the pressure difference reverses due to the high axial momentum of the inter-platform gap flow. Further downstream, the near-endwall flow turns toward the wedge face due to the cross-passage pressure gradient such that the PS flow crosses the inter-platform gap. The near-endwall flow gets ingested in the inter-platform gap such that a new boundary layer forms on the SS endwall, as shown in Fig. 6(c). Similar to the study by Lehmann et al. [13], the streamlines show no evidence of the passage vortex crossing the wedge face.

Fig. 7
Comparison of the endwall streamlines for every case in the simulation matrix
Fig. 7
Comparison of the endwall streamlines for every case in the simulation matrix
Close modal

The shroud endwall NuC contours displayed in Figs. 8(a)8(c) show that the flow ingestion in the LE region leads to high NuC on both platform edges. Around midchord, the vortex formed by the ejected gap flow creates a streak of low NuC on the SS endwall as it gets swept toward the suction surface. Further downstream, the restarting boundary layer leads to peak NuC on the SS platform edge, followed by a decrease in the streamwise direction, which is in agreement with the experimental results by Lehmann et al. [13].

Fig. 8
Comparison of the endwall NuC contours for every case in the simulation matrix
Fig. 8
Comparison of the endwall NuC contours for every case in the simulation matrix
Close modal

3.2 Backward-Facing Step in the TE Region.

The endwall surface streamlines in Fig. 7(e) show that an additional reattachment line, marked B, occurs downstream of reattachment line A on the SS endwall. This suggests that the flow separates at the PS step edge and reattaches on the SS platform downstream forming a recirculation zone in between. This resembles canonical backward-facing step flow as is sketched in Fig. 6(b). This hypothesis is supported by the reduced pressure margin in the aftchord region compared to an aligned endwall shown in Fig. 6(a).

The NuC contours displayed in Fig. 8(e) confirm the resemblance to canonical backward-facing step flow in the aftchord region, where a local maximum in NuC occurs along reattachment line B. At 75%Cx a reattachment NuC value of 3800 is obtained, which is an increase of 31% compared to the value obtained for an aligned endwall. The flow ingestion in the inter-platform gap leads to increased heat transfer in the recirculation zone compared to the canonical step flow without a gap since it accelerates the flow toward the step face. Due to this acceleration, the recirculating flow impinges on the PS wedge face, leading to a local NuC of 3800, which is more than six times the value observed for an aligned endwall.

Figures 7(d)7(f) show that the downstream reattachment length scales inversely to inter-platform gap width. This is likely to be caused by the ingestion rate, which controls the extent to which the reattaching streamline is pulled toward the wedge face. Since reattachment line B lies immediately downstream of the wedge face for the +1σ inter-platform gap, it is likely that for a further reduced step height to gap width ratio, the reattachment point moves to the SS wedge face, such that a new boundary layer forms on the SS endwall.

Figures 8(d)8(f) show that the peak NuC scales with the inter-platform gap width and thus the leakage flowrate, giving a peak NuC value of 2750 at 75%Cx for the 1σ gap and a value of 4900 for the +1σ gap. These correspond to a reduction of 28% and an increase of 29%, respectively, compared to the equivalent case with nominal gap.

Increasing the gap width also reduces the NuC value on the PS wedge face resulting in a value of 2450 for the +1σ gap, a reduction of 36% compared to the equivalent case with nominal gap. This is likely to be caused by the recirculating flow impinging on the PS wedge face with reduced intensity since most of this flow has been ingested in the gap before it reaches the wedge face.

3.3 Forward-Facing Step in the TE Region.

Figure 7(h) shows that the reattachment line B corresponding to canonical step flow reoccurs for a local forward-facing step. This is consistent with the static pressure increase at the wedge face compared to an aligned endwall, shown in Fig. 6(a), caused by the impingement of the upstream flow on the SS wedge face as shown in the canonical forward-facing step flow in Fig. 6(d).

Unlike for local backward-facing steps, reattachment line B bends toward the wedge face at the TE. The streamlines in the recirculation bubble are nearly parallel to the wedge face before getting ingested in the inter-platform gap. This suggests that part of the high axial momentum flow ejected from the inter-platform gap gets entrained in the recirculation bubble. Another reason for the more wedge face aligned flow in the recirculation bubble is the reduced ingestion of recirculating fluid into the inter-platform gap compared to a local backward-facing step since mainly the upstream PS platform flow gets ingested.

The NuC contours displayed in Fig. 8(h) show that NuC enhancement along reattachment line B is smaller than for the equivalent backward-facing step, with a NuC value of 2200, using the correction factor of 1.37, at 75%Cx. The largest NuC enhancement now occurs at reattachment line A, resulting in a maximum NuC value of 3100. This heat transfer increase is likely to be caused by the ejected gap flow now encountering a forward-facing step, which forces the flow to bend upward in order to emerge on the SS endwall. This upward movement leads to increased shear in the free shear layer forming at the SS platform edge and thus more energetic eddies impinging on the endwall at the reattachment point, leading to increased heat transfer.

The impingement of the PS endwall flow onto the SS wedge face leads to NuC values of up to 4000, more than double the expected value for an aligned endwall. Such increased heat transfer rates increase metal temperatures and increase the risk of oxidation and erosion on the wedge face.

The shroud endwall streamlines shown in Figs. 7(g)7(i) indicate that the reattachment length is less affected by a change in gap width compared to a local backward-facing step. This reinforces the hypothesis that the recirculating flow contained by reattachment line B is less influenced by the leakage flowrate.

Figures 8(g)8(i) show that for both considered gap width variations, the NuC levels at reattachment line B increase. For the 1σ gap width, the reattachment NuC value at 75%Cx is equal to 2660 using the correction factor of 1.37, which is an increase of 21% compared to the equivalent case with nominal gap. For the +1σ gap width, the peak NuC increases to a value of 2880 (using the correction factor of 1.37), which corresponds to a 31% increase.

The increase in NuC for a 1σ gap can be explained by the entrainment of the ejected gap flow in the recirculation bubble, which due to its high axial momentum increases the wall shear stress in the step-parallel direction. Figure 9 shows Cf,// contours on the shroud endwall for both a nominal and a 1σ inter-platform gap, where Cf,//=τ///12ρinVrel,in2. It can be observed that Cf,// increases by 35% at 75%Cx for the 1σ gap width compared to a nominal gap. Following the Reynolds analogy, this increases NuC in the aftchord region in addition to the increase in heat transfer caused by the reattaching flow.

Fig. 9
Comparison of Cf,// between a nominal (left) and a −1σ (right) gap for a 4σ FWD step
Fig. 9
Comparison of Cf,// between a nominal (left) and a −1σ (right) gap for a 4σ FWD step
Close modal

The maximum NuC value at reattachment line A decreases by 24% for a 1σ gap width, which is likely to be a result of the reduced ejected flowrate. For the increased inter-platform gap width, the peak NuC value changes by less than 10%. Also the peak NuC on the SS wedge face changes by less than 10%, for both considered gap width variations.

3.4 Strategy to Sentence Shroud Platform Steps.

The results of the 3D simulation matrix show that for an aligned endwall, the highest NuC values occur on the aftchord SS endwall edge where the boundary layer is restarting. For a 4σ backward-facing step, the highest NuC values are observed at the reattachment point on the aftchord SS endwall caused by the local canonical step flow. The reattachment NuC equally is a strong function of the inter-platform gap width. For the 4σ forward-facing step, the ejected gap flow reattaching on the SS endwall at midchord leads to the highest NuC enhancement.

The highest heat transfer enhancement has been observed on the aftchord region downstream of a local backward-facing step, which is therefore the most critical region from a part sentencing perspective, as this is where the part is most likely to get damaged. Since the flow is Q2D in the aftchord region, the flow parameters driving the change in heat transfer can be identified by conducting a parameter sweep on the Q2D shroud endwall step model for which both the domain and numerical methods are discussed in Sec. 2.2. These simulations are fast and easy to run, such that a simulation database is quickly acquired, but also allow to test the developed correlation on the HPTB shroud platform steps of the 3D simulation matrix using the same methods. However, since the correlation is ultimately used on manufactured parts, the final correlation is tuned using an experimental data base encompassing the required parameters, described in Sec. 6.

4 Shroud Step Nu Correlation Development

The parametrization of the shroud endwall step flow in the aftchord region uses a streamwise plane as shown in Fig. 1(c). The main parameter defining the flow field is the step height, which is included in the model as the characteristic length in the Reynolds number, ReH. Vogel and Eaton [24] indicated that also the step height to boundary layer thickness ratio influences the heat transfer at the reattachment point. Section 3 highlighted that the influence of the step on aerothermal performance changes depending on the local flow angle with respect to the step, α. Also the influence of the gap leakage flow, m˙L, on both the reattachment length and the reattachment Nu has been demonstrated. Finally, the studies by Awasthi et al. [25], and Zahn and Rist [26] suggest that increasing the step edge radius reduces the reattachment length for forward-facing steps.

The complete dimensionless parameter set is summarized in Table 3 including the reference values, where the value for H/θ corresponds to the step edge boundary layer thickness matching the step height. The leakage flowrate, which is determined by the inter-platform gap width and the local pressure gradient, is made dimensionless by the flow that is blocked by the step assuming a unit depth, m˙H, defined in Eq. (1).
(1)

The boundary layer thickness is represented by momentum thickness since it is an integrated property and therefore more robust to define for the nonuniform freestream occurring in the HPTB passage.

Table 3

Parameter set with reference case values

ReHαH/θm˙L/m˙Hr/H
760090 deg12.90.00.0
ReHαH/θm˙L/m˙Hr/H
760090 deg12.90.00.0

4.1 Parametric Study.

A database of Nusselt number profiles, with H as characteristic length, is created by conducting a parameter sweep on the Q2D shroud step domain for the variables listed in Table 3, using the numerical methods discussed in Sec. 2. Figure 10 shows the NuH profile along the step for the reference case described in Table 3. All NuH profiles have been normalized by the reference reattachment NuH value, NuH,ref, indicated by the blue dot in Fig. 10. Note that an equivalent reference case is used for the forward-facing step profiles.

Fig. 10
Illustration of the normalized NuH data presentation for the reference case
Fig. 10
Illustration of the normalized NuH data presentation for the reference case
Close modal
The NuH at the reattachment point is used as the performance parameter for the sentencing correlation. For every variable in Table 3, the relevant NuH profiles are shown and the corresponding reattachment NuH values are plotted as a function of the variable under study. A step NuH correlation is developed by fitting a curve through the reattachment NuH plot for each variable. The quality of the fitted curve is assessed using the normalized root mean square (RMSN) error given in Eq. (2), where the subscript Q2D represents a data point calculated using the Q2D model and the subscript corr represents the corresponding value given by the fitted curve.
(2)

The effect of each parameter is first considered in isolation. Since the shroud endwall step flow in the aftchord region is characterized by both a flow angle deviating from step-normal flow and inter-platform gap leakage, potential interactions between flow angle and leakage flow are studied afterward.

4.1.1 Isolated Effects.

The normalized NuH profiles and the corresponding reattachment NuH/NuH,ref values are shown in Fig. 11 for all parameters for a backward-facing step. It can be observed that the reattachment NuH scales almost linearly with ReH, with an RMSN error below 2%. In order to provide further validation of the used numerical methods and the obtained dependency on ReH, the Q2D model data points are displayed together with the results obtained using the 2D step heat transfer rig and the data from the study by Vogel and Eaton [24] in Fig. 12.

Fig. 11
NuH dependency on all parameters (BWD step)
Fig. 11
NuH dependency on all parameters (BWD step)
Close modal
Fig. 12
Comparison of the CFD results with both experimental data and the study by Vogel and Eaton [24] for different ReH
Fig. 12
Comparison of the CFD results with both experimental data and the study by Vogel and Eaton [24] for different ReH
Close modal

Figure 11 shows that both the peak NuH and the reattachment length decreases with the decreasing flow angle. However, the reduction in reattachment length in the considered flow angle range does not match the constant reattachment length observed by Kaltenbach [27] based on LES results. This is likely to be a consequence of the turbulence model being benchmarked using the step-normal flow, such that it is unable to capture this effect.

Even though the normalized reattachment NuH value is a strong function of flow angle, the decrease in NuH for a certain α is less than the decrease for the equivalent perpendicular flow with matched step-normal ReH. This implies that the reduction of the reattachment NuH with decreasing step-normal velocity component is counteracted by the increasing skin friction in the step-parallel direction, which following the Reynolds analogy increases the NuH of the step-parallel flow component.

Different heat transfer rates can be observed upstream of the step in the Nu profiles for varying step height to momentum thickness ratio in Fig. 11. This is caused by the variations in upstream boundary layer thickness used to change H/θ. When the reattachment NuH values are expressed as a function of H/θ, it can be observed that the peak NuH varies less than 10% in the considered H/θ range, indicating a weaker dependency compared to the previous two parameters. H/θ ratios up to 35 have been observed in the aftchord region of an HPT shroud endwall for the large 4σ step, indicating that the considered range is representative for more commonly occurring shroud platform steps.

The dependency of NuH on leakage flow rate is investigated using leakage flow rates representative of the aftchord region of an HPTB shroud endwall with a 4σ step, which ranges from 10% to 30% of m˙H. The NuH profiles in Fig. 11 show a decrease in reattachment length with the increased leakage flow ratio. The reattachment NuH scales approximately linear with the leakage flow ratio, where a leakage flow ratio equal to 13% leads to a 45% increase in peak NuH. The strong dependency of both the reattachment length and peak NuH on leakage flow ratio is in agreement with the observation made for a 3D shroud endwall in Sec. 3 and the result of the RANS study on 2D backward-facing steps with leakage flow ingestion by Abu-Nada et al. [28].

Figure 13 shows the NuH profiles and the corresponding peak values for all considered parameters for the Q2D forward-facing step model. The dependencies on ReH, α, and H/θ are nearly identical to the dependencies observed for backward-facing steps, whereas the increase in NuH for double the maximum leakage ratio considered for backward-facing steps is approximately 25%. The dependency on leakage flow ratio is thus weaker for forward-facing steps. This reduced sensitivity has also been observed in the aftchord region of an HPTB shroud endwall inSec. 3.

Fig. 13
NuH dependency on all parameters (FWD step)
Fig. 13
NuH dependency on all parameters (FWD step)
Close modal

The final row of Fig. 13 shows the NuH profiles for forward-facing steps with different step radii. It can be observed that the reattachment length reduces as the step radius increases, which is in agreement with the experimental results by Awasthi et al. [25] and the DNS results by Zahn and Rist [26]. Since assuming a constant NuH for different step edge radii leads to an RMSN error below 1.5%, no dependency on r/H is added to the correlation.

4.1.2 Interaction Between Flow Angle and Leakage.

Figure 14(a) shows the downstream NuH profiles for five different flow angles, with a leakage flow ratio of 8.9%. It can be observed that due to the reduced step-normal velocity for small α, the flow gets pulled back more strongly toward the wedge face by the inter-platform leakage, which results in an additional shortening of the reattachment length compared to the profiles shown in Fig. 11. The reattachment NuH is nearly constant for flow angles smaller than 45 deg, which implies that the effect of leakage on the reattachment NuH increases for small flow angles.

Fig. 14
Cross-effect between flow angle and leakage flowrate on the reattachment NuH (BWD step): (a) NuH profiles for m˙L/m˙H=8.9% and (b) reattachment NuH values
Fig. 14
Cross-effect between flow angle and leakage flowrate on the reattachment NuH (BWD step): (a) NuH profiles for m˙L/m˙H=8.9% and (b) reattachment NuH values
Close modal

Figure 14(b) shows the peak NuH as a function of both leakage flow rate and angle with respect to the step. It can be observed that for leakage flow ratios exceeding 8.8%, the reattachment NuH increases with the decreasing flow angle. The peak NuH asymptotes when the angle reduces toward the step-parallel flow. This illustrates that the flow mechanism changes from reattachment on the downstream platform to reattachment in the gap such that a new boundary layer starts on the downstream surface. The flow angle at which this change in flow reattachment occurs depends on the leakage flowrate, such that the dependency of peak NuH on flow angle can be described using a blending function based on the leakage flowrate. Figure 14(b) shows that this blending function limits the RMSN error to 1.8% for m˙L/m˙H equal to 13.3%.

Figure 15(a) shows that the interaction between leakage flowrate and flow angle is weaker for a forward-facing step. This can be attributed to the fact that mainly upstream platform flow gets ingested in the gap instead of the recirculating fluid, as has been observed in Sec. 3. However, for flow angles below 60 deg, a disproportional increase in NuH can still be noted, such that the RMSN error would increase to 3.5% for a leakage mass flow ratio of 26.6%. In order to account for this disproportional enhancement, a flow angle dependent factor is added to the leakage term, which increases the NuH at small flow angles. Figure 15(b) shows that this additional factor reduces the RMSN error to 1.6%.

Fig. 15
Cross-effect between flow angle and leakage flowrate on the reattachment NuH (FWD step): (a) NuH profiles for m˙L/m˙H=8.9% and (b) reattachment NuH values
Fig. 15
Cross-effect between flow angle and leakage flowrate on the reattachment NuH (FWD step): (a) NuH profiles for m˙L/m˙H=8.9% and (b) reattachment NuH values
Close modal

4.2 Key Parameters Influencing NuH.

It can be concluded that the reattachment NuH is mainly dependent on the step Reynolds number, flow angle, and leakage flow ratio. This results in a step Nu sentencing correlation of the form
(3)
where the dependencies on the input parameters vary between backward- and forward-facing steps.

5 Testing the Correlation on HPTB Shroud Endwall Steps

In order to assess the step correlation performance on the geometry it is meant to sentence, it is tested on all the aftchord shroud endwall steps discussed in Sec. 3. In order to minimize the additional resources required to use the step sentencing correlation on the production line, the input variables are calculated from a design-intent computational fluid dynamics (CFD) simulation and a step height measurement. Therefore, the input variables for these test cases are extracted on the shroud endwall from the aligned platform case with nominal gap. The detailed calculation of the input parameters is explained in the  Appendix.

The step correlation is tested on four streamwise slices on the aftchord shroud endwall, labeled A–D in Fig. 16. The extraction of the two-run NuH to which the prediction is compared is shown for streamwise slice C. Note that the reattachment length has decreased from 6.5H to 2.5H due to the reduced flow angle relative to the step edge and the flow ingestion in the gap.

Fig. 16
Extraction of the NuH value on the endwall
Fig. 16
Extraction of the NuH value on the endwall
Close modal

Figure 17 shows the NuH prediction accuracy for the four streamwise slices for all step cases in the 3D simulation matrix. The influence of gap leakage flow for backward-facing steps is illustrated by the NuH increase for all slices on a test case with the increasing gap width. Such a division in clusters cannot be observed for forward-facing steps.

Fig. 17
Q2D shroud step correlation performance
Fig. 17
Q2D shroud step correlation performance
Close modal

All predictions for backward-facing steps are within 20%, where the highest relative error is observed for plane A in the case of a +1σ gap. In this case, the flow is close to the changing point between reattaching backward-facing step flow and a restarting boundary layer downstream, where the correlation is sensitive to slight changes in the input parameters. Also the forward-facing step predictions are within 20%, except for the point representing plane A for a 1σ gap, which lies closest to midchord where the ejected gap flow gets entrained in the recirculation bubble. It has been shown in Sec. 3 that due to the large axial momentum of the entrained flow and the reduced ingestion into the inter-platform gap, the step-parallel component of wall shear stress increases over the expected value for an equivalent Q2D forward-facing step. Nevertheless, the prediction error remains below 30%, which emphasizes the robustness of the correlation to deviations from the Q2D flow.

6 Tuning Using Experimental Data

In this section, the shroud sentencing correlation developed using a numerical study on the Q2D step model is tuned using experimental results. This is done to mitigate the limitations of using steady RANS CFD for the physics associated with canonical backward- and forward-facing step flow, which are inherently unsteady flows due to the vortex shedding at the step edge [29,30]. Moreover, vortices comparable in size to the step height occur in the free shear layer formed between the freestream and the recirculation zone at the step edge, resulting in anisotropic turbulence, which is incompatible with the Boussinesq approximation [29]. However, the 2D backward-facing step has been used to tune turbulence models to improve their performance on turbulent separation [29,31,32]. Menter [33] showed that this has led to improved predictions of both reattachment length and shear stress downstream of the step using the kω model, which is the near-wall model of the kω SST model used in this study.

First, correction factors for the RANS prediction of the reattachment Nu are obtained by comparing the simulations results on the 2D step heat transfer rig model with both backward- and forward-facing steps to the experimental data. Afterwards, the dependencies in the step heat transfer correlation obtained using the Q2D shroud step model are tuned using experimental data.

6.1 Comparison With the Step Heat Transfer Rig Model.

The comparison of the NuH profile along the middle of a backward-facing step between CFD and experiment is shown in Fig. 18(a), both as a function of Rex and X/H. The zero pressure gradient (ZPG) Nux correlations taken from the review paper by Lienhard [34] are plotted to provide a reference, where the turbulent correlation is moved to the location of the boundary layer trip in the experiment. The NuH profiles upstream of the step edge lie within 5% of each other and the ZPG turbulent correlation, which is within the experimental accuracy. The slope of the NuH profiles match in the turbulent region upstream of the step, showing that the boundary layer growth is matched.

Fig. 18
NuH comparison between CFD and experiment: (a) backward-facing step and (b) forward-facing step
Fig. 18
NuH comparison between CFD and experiment: (a) backward-facing step and (b) forward-facing step
Close modal

In order to compare the difference in NuH at the reattachment point, the results are displayed as a function of X/H and are both aligned and normalized using the NuH value 20H upstream of the step. The NuH at the reattachment point is underpredicted by only 3%, whereas the reattachment length is overpredicted by 2H. This shows that benchmarking turbulence models on a backward-facing step has led to predictions within 5% despite the occurring flow physics not being modeled correctly.

Figure 18(b) shows the comparison between the experimental and numerical NuH profile for a forward-facing step. Again, the NuH decay upstream of the step matches and both profiles lie within 10% of the turbulent ZPG correlation. The NuH decay is steeper in this case since the upstream near-wall flow encounters an adverse pressure gradient created by the static pressure maximum at the step face. Immediately upstream of the step, the numerical NuH prediction drops because the CFD does not capture the unsteady growing and shrinking of the upstream recirculation zone [35] but models it as a steady recirculation zone.

When the normalized NuH are displayed as a function of X/H, it can be observed that the experimentally measured reattachment point lies approximately 3H downstream of the step edge, which is in agreement with the literature [30,36]. The CFD, on the other hand, predicts a reattachment length of 7.2H. The corresponding reattachment NuH value is underpredicted by 37% in the CFD. Therefore, a correction factor of 1.37 is used on forward-facing step cases in the 3D simulation matrix in Sec. 3.

6.2 Tuning of the Shroud Step Correlation.

The dependencies of the shroud step correlation developed using the Q2D simulations are tuned by comparing it to experimental reattachment NuH values. In order to remove the effect of the non-engine representative H/S in the rig, the correlation output for the reference case is scaled to match the corresponding experimental value. In order to further facilitate comparing the experimental data, the NuH profiles are aligned 5H upstream of the step, where the boundary layer, and thus NuH, is expected to be identical for all cases. Only the NuH profiles for different ReH are not scaled since the upstream boundary layer is affected by the change in flow conditions.

6.2.1 Backward-Facing Steps.

The measured NuH profiles and the tuning of the correlation for all variables are shown in Fig. 19 for a backward-facing step. The experimental uncertainty of 5% is indicated by the vertical error bars, and the uncertainty of the measured leakage flowrate is indicated by a horizontal error bar where applicable. The Nu profiles for different α in Fig. 19(a) show that the reattachment length is constant, which is in agreement with the study by Kaltenbach [27] and confirm that RANS simulations erroneously predict a shortening of the reattachment length in the considered range of flow angle. In the case of leakage flowrate, Fig. 19(a) shows that a leakage flow ratio of 9% reduces the reattachment length by 0.75H, which confirms the trend observed in both the 3D shroud endwall and the Q2D database simulations. The corresponding heat transfer increase is equal to 15.6%, which is underpredicted by the correlation and the dependency is tuned as shown by the black line in Fig. 19(b). For all other variables, including the insensitivity to step edge radius, the experimental data lie close to the trends in the correlation and no tuning is required.

Fig. 19
Experimental NuH profiles and tuning of the correlation (BWD): (a) NuH profiles and (b) tuning of correlation
Fig. 19
Experimental NuH profiles and tuning of the correlation (BWD): (a) NuH profiles and (b) tuning of correlation
Close modal

6.2.2 Forward-Facing Steps.

The NuH profiles and the tuning of the shroud step correlation are shown for all parameters in Fig. 20 for a forward-facing step. Figure 20(a) shows that for varying flow angle, the reattachment length only shortens when α reduces from 60 deg to 45 deg. This behavior is not captured by RANS simulations, which predict a continuously reducing reattachment length. For the varying leakage flow ratio, the reattachment length reduces with 0.83H for a leakage flow ratio of 23.2%, which is similar to the decrease observed for backward-facing steps for a leakage flow ratio of 9.0%. This confirms the reduced dependency of the reattachment length on leakage flow rate for forward-facing steps observed in Sec. 3. Finally, the reducing reattachment length with increasing step edge radius can be observed and is in agreement with the Q2D CFD results and the available literature [25,26].

Fig. 20
Experimental NuH profiles and tuning of the correlation (FWD): (a) NuH profiles and (b) tuning of correlation
Fig. 20
Experimental NuH profiles and tuning of the correlation (FWD): (a) NuH profiles and (b) tuning of correlation
Close modal

Figure 20(b) shows that the unmatched experimental ReH data point exceeds the value predicted by the correlation. Therefore, the dependency on ReH is adjusted as shown by the black line. The accuracy of this correction decreases for ReH values outside of the two measured data points. However, since the ReH for HPTB shroud platform steps representative of manufacturing variability is unlikely to exceed 15,000, this mainly risks overpredicting the reattachment NuH value when ReH<7600.

In the case of varying α, the experimental reattachment NuH closely matches the correlation from 90 deg to 60 deg followed by a deviation as the reattachment NuH remains constant between 60 deg and 45 deg. Since this change in trend is unexpected, it is likely that the actual NuH for a flow angle of 45 deg lies toward the lower end of the error bar and the deviation from the correlation is small.

Finally, Fig. 20(b) shows that the Q2D CFD underpredicts the dependency of the reattachment NuH on leakage flow rate and does not predict the increased reattachment NuH with increasing step edge radius. Therefore, the correlation is tuned as shown by the black lines. Note that since platform edges tend to round off due to wear, the dependency on step edge radius suggests that the heat transfer enhancement caused by the step increases with time in service.

7 Conclusions

This study showed that a correlation based on a parametrized Q2D model of a shroud endwall step is able to predict the heat transfer enhancement caused by aftchord shroud platform steps with an accuracy of 20% for the majority of cases. This follows from the fact that the shroud platform flow is Q2D in the aftchord region, both with and without platform steps, as has been shown by a steady RANS study of engine-representative shroud platform manufacturing variations in a three passage HPT rotor model. The shroud step heat transfer correlation has been tuned using experimental data such that it is ready to triage manufactured HPTB on the production line.

Acknowledgment

The authors gratefully acknowledge InnovateUK via the Aerospace Technologies Institute for the numerical and experimental resources. The authors gratefully acknowledge Rolls-Royce for permission to publish. The first author was funded by a joint Rolls-Royce and EPSRC Ph.D. studentship.

Conflict of Interest

There are no conflicts of interest.

Data Availability Statement

The datasets generated and supporting the findings of this article are obtainable from the corresponding author upon reasonable request.

Appendix: Calculating the Correlation Input Variables

The input variable extraction for the four planes on an HPTB endwall discussed in the article is illustrated in Fig. 21. The data extraction plane is parallel to the wedge face but is moved 1.25%H upstream in the direction perpendicular to the step. This is done in order to avoid the local acceleration caused by the leakage flow influencing the measured boundary layer and flow angle profile. The four wall-normal lines in Fig. 21 show where the four streamwise test planes intersect the data extraction plane. On each of these slices, the wall-normal profile of flow angle relative to the step, velocity, density, and dynamic viscosity is extracted.

Fig. 21
Illustration of the input data extraction
Fig. 21
Illustration of the input data extraction
Close modal
The momentum thickness is calculated iteratively using Eq. (A1), where the upper limit of the integration starts at the first data point and proceeds in the wall-normal direction. The momentum thickness is defined at the point where the relative change in momentum thickness is below 0.75% and the boundary layer thickness is defined as the integration limit at that point. This allows to use Eq. (A2) to calculate the flow angle and the values at the boundary layer edge to calculate ReH.
(A1)
(A2)

Unfortunately, the leakage flow rate cannot be extracted from the aligned reference case since the driving pressure difference is dependent on both the local step and the inter-platform gap width. Therefore, the leakage flowrate is extracted using a local mass flow monitor positioned in the gap for every streamwise slice in the respective shroud platform step simulation. Since the leakage flow needs to be calculated assuming a unit width, the leakage flow is normalized by the length of the monitor in the direction parallel to the step.

Nomenclature

p =

pressure (Pa)

r =

radius (m)

w =

gap width (m)

C =

chord (m)

H =

step height (m)

S =

span (m)

T =

temperature (K)

V =

velocity (ms1)

X =

streamwise coordinate (m)

m˙H =

=ρVH, step blocked mass flow (kg/s)

m˙L =

inter-platform gap leakage mass flow (kg/s)

Cx =

axial chord (m)

y+ =

dimensionless wall distance

Tu =

turbulence intensity (%)

Greek Symbols

α =

flow angle (deg)

ϵ =

error

θ =

momentum thickness (m)

μ =

mean

ρ =

density (kg/m3)

σ =

standard deviation

τ =

shear stress (Nm/2)

Dimensionless Groups

Cp =

pressure coefficient

Cf =

friction coefficient

 Nu =

Nusselt number

Re =

Reynolds number

Superscripts and Subscripts

aw =

adiabatic wall

 corr =

correlation value

in =

inlet value

is =

isentropic value

ref =

reference value

rel =

relative value

w =

wall

=

freestream value

// =

step-parallel value

Abbreviations

BWD =

backward-facing

CFD =

computational fluid dynamics

DNS =

direct numerical simulation

FWD =

forward-facing

HPTB =

high-pressure turbine blade

HTC =

heat transfer coefficient

IR =

infrared

LE =

leading edge

LES =

large eddy Simulation

PDF =

probability density function

PS =

pressure side

Q2D =

quasi-2D

RANS =

Reynolds-averaged Navier–Stokes

RMSN =

root mean square normalized

SS =

suction side

SST =

shear stress transport

STL =

StereoLithography

TE =

trailing edge

ZPG =

zero pressure gradient

References

1.
Ackert
,
S.
,
2011
, “
Engine Maintenance Concepts for Financiers
,”
Aircraft Monitor
,
2
, pp.
1
43
.
2.
Grewe
,
R. P.
,
Miller
,
R. J.
, and
Hodson
,
H. P.
,
2014
, “
The Effect of Endwall Manufacturing Variations on Turbine Performance
,”
Proceedings of the ASME Turbo Expo 2014: Turbine Technical Conference and Exposition
,
Düsseldorf, Germany
,
June 16–20
,
p. V02CT38A008, ASME Paper No. GT2014-25326
.
3.
Lee
,
W. Y.
,
Dawes
,
W. N.
, and
Coull
,
J. D.
,
2020
, “
Physics-Based Part Orientation and Sentencing: A Solution to Manufacturing Variability
,”
ASME J. Turbomach.
,
142
(
10
), p.
101001
.
4.
Cardwell
,
N. D.
,
Sundaram
,
N.
, and
Thole
,
K. A.
,
2006
, “
Effect of Midpassage Gap, Endwall Misalignment, and Roughness on Endwall Film-Cooling
,”
ASME J. Turbomach.
,
128
(
1
), pp.
62
70
.
5.
Heinze
,
K.
,
Meyer
,
M.
,
Scharfenstein
,
J.
,
Voigt
,
M.
, and
Vogeler
,
K.
,
2014
, “
A Parametric Model for Probabilistic Analysis of Turbine Blades Considering Real Geometric Effects
,”
CEAS Aeronaut. J.
,
5
(
1
), pp.
41
51
.
6.
Högner
,
L.
,
Nasuf
,
A.
,
Voigt
,
P.
,
Voigt
,
M.
,
Vogeler
,
K.
,
Meyer
,
M.
,
Berridge
,
C.
, and
Goenaga
,
F.
,
2016
, “
Analysis of High Pressure Turbine Nozzle Guide Vanes Considering Geometric Variations
,”
Proceedings of ASME Turbo Expo 2016: Turbomachinery Technical Conference and Exposition
,
Seoul, South Korea
,
June 13–17
,
p. V02CT45A025, Paper No. GT2016-57502
.
7.
Voigt
,
P.
,
Högner
,
L.
,
Fiedler
,
B.
,
Voigt
,
M.
,
Mailach
,
R.
,
Meyer
,
M.
, and
Nasuf
,
A.
,
2019
, “
Comprehensive Geometric Description of Manufacturing Scatter of High-Pressure Turbine Nozzle Guide Vanes for Probabilistic CFD Analysis
,”
ASME J. Turbomach.
,
141
(
8
), p.
081002
.
8.
Garzón
,
V. E.
, and
Darmofal
,
D. L.
,
2003
, “
Impact of Geometric Variability on Axial Compressor Performance
,”
ASME J. Turbomach.
,
125
(
4
), pp.
692
703
.
9.
Duffner
,
J. D.
,
2008
, “The Effects of Manufacturing Variability on Turbine Vane Performance,” Master’s thesis, Massachusetts Institute of Technology, Cambridge, Boston, MA.
10.
Hada
,
S.
, and
Thole
,
K. A.
,
2011
, “
Computational Study of a Midpassage Gap and Upstream Slot on Vane Endwall Film-Cooling
,”
ASME J. Turbomach.
,
133
(
1
), p.
011024
.
11.
Lange
,
E.
,
Lynch
,
S.
, and
Lewis
,
S.
,
2016
, “
Computational and Experimental Studies of Midpassage Gap Leakage and Misalignment for a Non-Axisymmetric Contoured Turbine Blade Endwall
,”
Proceedings of ASME Turbo Expo 2016: Turbomachinery Technical Conference and Exposition
,
Seoul, South Korea
,
June 13–17
,
p. V05AT13A001, Paper No. GT2016-56128
.
12.
Zhang
,
K.
,
Li
,
Z.
, and
Li
,
J.
,
2019
, “
Effect of Gas Turbine Endwall Misalignment With Slashface Leakage on the Blade Endwall Aerothermal Performance
,”
Int. J. Heat. Mass. Transfer.
,
131
(
1
), pp.
1247
1259
.
13.
Lehmann
,
K.
,
Kanjirakkad
,
V.
, and
Hodson
,
H.
,
2012
, “
Aerodynamic and Aerothermal Investigation of the Flow Around an HPT Rotor Shroud: Heat Transfer and Cooling Effectiveness
,”
Proceedings of the ASME 2011 Turbo Expo: Power for Land, Sea, and Air
,
British Columbia, Canada
,
June 6–10
, pp.
1375
1384
.
14.
Rushton
,
G. J.
,
2003
, “High-Pressure Turbine Shroud Leakage,” Ph.D. thesis, University of Cambridge, Cambridge, UK.
15.
Wallis
,
A. M.
,
Denton
,
J. D.
, and
Demargne
,
A. A. J.
,
2000
, “
The Control of Shroud Leakage Flows to Reduce Aerodynamic Losses in a Low Aspect Ratio, Shrouded Axial Flow Turbine
,”
ASME J. Turbomach.
,
123
(
2
), pp.
334
341
.
16.
Lehmann
,
K.
,
Thomas
,
R.
,
Hodson
,
H.
, and
Stefanis
,
V.
,
2010
, “
Heat Transfer and Aerodynamics of Over-Shroud Leakage Flows in a High-Pressure Turbine
,”
Proceedings of the ASME Turbo Expo 2009: Power for Land, Sea, and Air
,
Orlando, FL
,
June 8–12
, pp.
515
525
.
17.
Lee
,
W. Y.
,
Dawes
,
W. N.
,
Coull
,
J. D.
, and
Goenaga
,
F.
,
2019
, “
The Impact of Manufacturing Variability on Multi-Passage High Pressure Turbine Aerodynamics
,”
AIAA Scitech 2019 Forum
,
San Diego, CA
,
Jan. 7–11
,
p. 1950
.
18.
Braslow
,
A. L.
,
Hicks
,
R. M.
, and
Harris
,
R. V.
,
1966
, “Use of Grit-Type Boundary-Layer-Transition Trips on Wind-Tunnel Models,”
National Aeronautics and Space Administration Langley Field
,
VA
, Technical Report No. NASA TN D-3579.
19.
Oldfield
,
M. L. G.
,
2008
, “
Impulse Response Processing of Transient Heat Transfer Gauge Signals
,”
ASME J. Turbomach.
,
130
(
2
), p.
021023
.
20.
Collins
,
M.
,
Chana
,
K.
, and
Povey
,
T.
,
2016
, “
Improved Methodologies for Time-Resolved Heat Transfer Measurements, Demonstrated on an Unshrouded Transonic Turbine Casing
,”
ASME J. Turbomach.
,
138
(
11
), p.
111007
.
21.
Playford
,
W.
,
2018
, “Well-Conditioned Heat Transfer Measurements on Engine Scale Gas Turbine Rigs,” Ph.D. thesis, University of Cambridge, Cambridge, UK.
22.
Tucker
,
P. G.
,
2016
,
Advanced Computational Fluid and Aerodynamics
(
Cambridge Aerospace Series
),
Cambridge University Press
,
Cambridge, UK
.
23.
Wang
,
T.
,
Zhao
,
Y.
,
Leggett
,
J.
, and
Sandberg
,
R. D.
,
2023
, “
Direct Numerical Simulation of a High-Pressure Turbine Stage: Unsteady Boundary Layer Transition and the Resulting Flow Structures
,”
ASME J. Turbomach.
,
145
(
12
), p.
121009
.
24.
Vogel
,
J. C.
, and
Eaton
,
J. K.
,
1985
, “
Combined Heat Transfer and Fluid Dynamic Measurements Downstream of a Backward-Facing Step
,”
ASME. J. Heat. Transfer-Trans.
,
107
(
4
), pp.
922
929
.
25.
Awasthi
,
M.
,
Devenport
,
W. J.
,
Alexander
,
W. N.
, and
Glegg
,
S. A. L.
,
2019
, “
Aeroacoustics of Rounded Forward-Facing Steps: Near-Field Behavior
,”
AIAA. J.
,
57
(
3
), pp.
1237
1249
.
26.
Zahn
,
J.
, and
Rist
,
U.
,
2017
, “
Active and Natural Suction at Forward-Facing Steps for Delaying Laminar–Turbulent Transition
,”
AIAA. J.
,
55
(
4
), pp.
1343
1354
.
27.
Kaltenbach
,
H.-J.
,
2004
, “
Turbulent Flow Over a Swept Backward-Facing Step
,”
Eur. J. Mech. - B/Fluids
,
23
(
3
), pp.
501
518
.
28.
Abu-Nada
,
E.
,
Al-Sarkhi
,
A.
,
Akash
,
B.
, and
Al-Hinti
,
I.
,
2007
, “
Heat Transfer and Fluid Flow Characteristics of Separated Flows Encountered in a Backward-Facing Step Under the Effect of Suction and Blowing
,”
ASME. J. Heat. Transfer-Trans.
,
129
(
11
), pp.
1517
1528
.
29.
Eaton
,
J. K.
, and
Johnston
,
J. P.
,
1981
, “
A Review of Research on Subsonic Turbulent Flow Reattachment
,”
AIAA. J.
,
19
(
9
), pp.
1093
1100
.
30.
Sherry
,
M.
,
Lo Jacono
,
D.
, and
Sheridan
,
J.
,
2010
, “
An Experimental Investigation of the Recirculation Zone Formed Downstream of a Forward Facing Step
,”
J. Wind Eng. Ind. Aerodyn.
,
98
(
12
), pp.
888
894
.
31.
Kim
,
J.
,
Ghajar
,
A. J.
,
Tang
,
C.
, and
Foutch
,
G. L.
,
2005
, “
Comparison of Near-Wall Treatment Methods for High Reynolds Number Backward-Facing Step Flow
,”
Int. J. Comput. Fluid Dyn.
,
19
(
7
), pp.
493
500
.
32.
Thangam
,
S.
, and
Speziale
,
C. G.
,
1992
, “
Turbulent Flow Past a Backward-Facing Step—A Critical Evaluation of Two-Equation Models
,”
AIAA. J.
,
30
(
5
), pp.
1314
1320
.
33.
Menter
,
F. R.
,
1994
, “
Two-Equation Eddy-Viscosity Turbulence Models for Engineering Applications
,”
AIAA. J.
,
32
(
8
), pp.
1598
1605
.
34.
Lienhard
,
J. H., V
,
2020
, “
Heat Transfer in Flat-Plate Boundary Layers: A Correlation for Laminar, Transitional, and Turbulent Flow
,”
ASME. J. Heat. Transfer-Trans.
,
142
(
6
), p.
061805
.
35.
Pearson
,
D. S.
,
Goulart
,
P. J.
, and
Ganapathisubramani
,
B.
,
2013
, “
Turbulent Separation Upstream of a Forward-Facing Step
,”
J. Fluid. Mech.
,
724
, pp.
284
304
.
36.
Leclercq
,
D.
,
Jacob
,
M.
,
Louisot
,
A.
, and
Talotte
,
C.
,
2001
, “
Forward-Backward Facing Step Pair—Aerodynamic Flow, Wall Pressure and Acoustic Characterisation
,”
7th AIAA/CEAS Aeroacoustics Conference and Exhibit 2001
,
Maastricht, The Netherlands
,
May 28–30
,
p. 2249
.