Abstract

For novel high-speed small core turbines, with tip clearance below 0.5 mm, even small blade-to-blade tip clearance variation is significant. The assessment of these complex flows is pertinent to the design of the next generation of small-core turbines. This paper provides a thorough experimental analysis of the shroud's unsteady heat flux and static pressure in a small-core squealer-tip blade turbine. Atomic layer thermopile sensors (ALTPs) and fast response pressure transducers were used to perform high-frequency acquisition at 2 MHz around the 51% axial blade chord on the shroud. Measurements were taken at engine-representative conditions at several operational conditions and tip clearances. The signals were phase-locked averaged (PLA) over the revolution period and synchronized to identify individual blade and row signature. The linear relationship between the rotational tip Reynolds and the static pressure ratio across the blade tip reveals the transition point to reverse over-tip flows. Total heat flux is decomposed into the different steady and unsteady heat flux contributions. It is demonstrated that the adiabatic wall temperature governs the unsteady heat flux and contributes to one-third of the total surface heat flux. A linear trend was observed between the unsteady heat flux and the tip clearance measured at the pressure and suction side rims. Similar trends were observed between the local heat flux and the pressure ratio across the tip. A comparison with computational fluid dynamics (CFD) predictions highlights some limitations on resolving the detached and secondary flows, evidencing the necessity of complementary experimental data.

1 Introduction

Tip leakage flows in unshrouded turbines lead to high aerothermal penalties in the machine's performance. Pressure gradients across the blade tip drive hot flow from the main passage toward the tip region, increasing the heat loads in the blade tip and the over-tip shroud [1,2]. Within the blade tip region, additional thermal loads arise from air-to-surface friction [3] and pressure losses result from the interaction of the over-tip flows with pre-existing flow structures [4,5]. Eventually, the tip leakage flows exit the tip region, generating and feeding into the tip leakage vortex [6]. This represents another source of entropy due to viscous friction with the main passage flow [3]. Additionally, the development of this tip leakage vortex further influences the exit flow angle and, consequently, impacts the performance of downstream stages [7]. The tip clearance size plays a significant role in the formation and amount of leakage flow [811], and therefore influences the magnitude of these phenomena. These penalties become extremely relevant for small-core turbomachinery that feature short flow passage heights and low aspect ratios of blades and vanes. In this novel design configuration, the growth of any vortex toward the main passage generates higher levels of blockage that hinder the turbine efficiency [12]. Small core turbines with very tight clearances below 0.5 mm operate at higher rotational speeds [11,13]. Therefore, the shroud is subject to larger unsteady frequencies and small blade-to-blade tip clearance variations can have substantial effects on the machine [14].

To evaluate a design's performance under various tip clearance conditions, computational tools are often used to carry out unsteady flow simulations. The elevated difficulties and computational costs associated with performing high fidelity direct numerical simulation simulations are often grounds to perform lower fidelity Reynolds-averaged Navier-Stokes (RANS) simulations instead. However, its limitations in resolving unsteady detached and secondary flows in very tight clearance configurations [15,16], and the need for computational tool validation, justify the necessity of complementary experimental studies.

To identify the work processes taking place around the tip clearance region, several studies analyzed the adiabatic wall temperature and heat transfer coefficient. Thorpe et al. [17] related the decrease in temperature with the work extraction processes through the rotor passage. Their results revealed that the blade tip can perform work on the flow, opposing the work extraction of the rotor row, and increase the temperature above the turbine inlet total temperature. While some studies, such as those performed by Wheeler [18,19] and Krishnababu [20,21], demonstrate a significant influence between the clearance size and the aerothermal performance, other research only associated the variations in clearance with the aerodynamic signatures. Lavagnoli et al. [22] studied the Nusselt number and pressure field fluctuations and only the latter showed a direct correlation with tip clearance signatures. A slight match was observed with the adiabatic wall temperature only during spinning-up processes, indicating that blade-to-blade variability was likely generated by transient mechanisms only.

While past results were obtained in large scale, high aspect ratio turbine rigs, this paper presents high frequency measurements of the instantaneous tip clearance, unsteady shroud static pressure and heat flux of a small-core squealer-blade tip turbine configuration. The measurements are taken at engine-representative conditions at various pressure ratios and tip clearance sizes. The understanding of the flow field generated by the passing blades is obtained from the phase-resolved evaluation of the different parameters, and governing relationships are observed from the evaluation of the various blade row signatures.

2 Experimental Setup

2.1 Facility.

The tests were performed in the pressure-driven wind tunnels of Purdue Experimental Turbine Aerothermal Lab (PETAL) [23], at Purdue University, schematized in Fig. 1. It features a storage capacity of 56 m3 of compressed dry air at 150 bar (∼2000 psi). Several lines exit from this supply to feed the experimental apparatus. A first supply line provides unheated air directly from the storage tanks. A second line passes through a heat exchanger within a natural gas air heater capable of heating the air to 1500 ′F. The hot and cold supplies are then mixed to obtain the desired air temperature conditions. The mixed flow goes through a calibrated critical venturi that provides an accurate measurement of the air mass flow. Downstream of the venturi, a set of fast actuation butterfly valves direct the flow either through a purge line discharging to the atmosphere, or toward the test section. Immediately upstream of the test section, the flow is conditioned in a settling chamber and equipped with a series of screens and honeycombs to deliver clean, homogeneous, and uniform flow conditions. Downstream of the test section, the air goes through a sonic valve. The flow area of the sonic valve can be adjusted, allowing the operator to set the downstream pressure of the test section, thus ensuring a steady pressure ratio across the test section in subsonic conditions.

Fig. 1
Schematic of the Purdue PETAL blowdown facility
Fig. 1
Schematic of the Purdue PETAL blowdown facility
Close modal

The facility can be operated both in transient and steady modes at pressure ratios across the test section up to 6. The steady operation of the facility for more than 15 min allows studying the thermal equilibrium and the effect of thermal expansion on tip clearance size and, consequently, the effect of varying tip clearance at constant turbine operational conditions.

2.2 Test Section.

The empirical measurements were taken in the STARR test section (Small Turbine Aerothermal Rotating Rig), portrayed in Fig. 2. The STARR rig [24] is a small-core, two-stage high-speed turbine rig. The blade rows were equipped with squealer tips featuring a continuous rim all around the blade tip edge, as exemplified in the same figure. The aft end of the turbine shaft extends through the exhaust plenum and sonic valve and connects directly to a high-speed synchronous electric motor. The dyno has a motoring power of up to 100 HP, which is used to spin the turbine up to 3000 rpm to warm the bearings prior to flowing air. Once flow is established through the test section, the dyno has an absorption capability of 1 MW, enabling continuous loaded testing at speeds above 14,000 rpm.

Fig. 2
Starr test section, high frequency instrumentation, and measurement locations
Fig. 2
Starr test section, high frequency instrumentation, and measurement locations
Close modal

The coupling of the STARR test section with the electric dyno allows setting different rotational speeds, and therefore tip clearance sizes, throughout the entire span of pressure ratios of the facility.

2.3 Over-Tip Instrumentation.

Several high frequency sensors were installed in a series of inserts mounted to the over-tip shroud at different circumferential locations as illustrated in Fig. 2. The inserts follow the shroud curvature and include an O-ring to prevent leakage around the insert. All the over-tip measurements were performed at an axial location equal to x/c=0.51, c being the axial blade chord length.

2.3.1 Tip Clearance Measurements.

Tip clearance was measured with Pseudo-Triaxial capacitance probes with a 2.5 mm electrode. The Fogale Capablade Fusion System [25] was used as the conditioning and acquisition system. This mainframe was equipped with six MC925 translational modules [26] and applied a 30 V/pF gain to the probes. All the acquisition was continuously performed at a sample frequency of 2 MHz, and the sensor calibrations were performed following the methodologies explained in Refs. [27] and [28].

2.3.2 Unsteady Pressure Measurements.

The unsteady static pressure at the over-tip shroud was acquired with Kulite miniature pressure transducers. Sensor models XCE-062-100A and XCE-062-50A [29] were used on stage 1 and 2, respectively. A Precision Filter Kulite Conditioner model 10,177 [30] output independent static and dynamic signals for each input Kulite sensor. These are obtained by applying a Butterworth low-pass filter with f3dB=322Hz and a high-pass filter of the same type with f3dB=160Hz to each input Kulite signal. The outputs of the conditioning system were recorded by a Genesis GEN2tB Transient Recorded and Data Acquisition System [31]. This was equipped with two GN8101B cards [32] for independent low and high frequency acquisitions. All the measurements were performed at 1 MHz, applying an additional digital low-pass filter with f3dB=250KHz to both readings.

2.3.3 Unsteady Heat Flux Measurements.

The unsteady heat flux was measured with an Atomic Layer Thermopile (ALTP) [33]. The sensor sensitivity was 31 uV/W/cm2, and an amplifier provided a ×200 gain to the sensor output. A Picoscope 5442D MSO [34] recorded the amplified signal at a rate of 2 MHz without any additional filtering.

2.4 Tachometer Signal.

For blade identification and tracking, all the measurements were triggered by a once-per-revolution signal provided by a Shimpo MCS-655M Retro-reflective Speed sensor [35]. This sensor was installed at top dead center (TDC) (0 deg ALF circumferential location) facing a photoreflective tape in the shaft, providing a square wave between 5 V and 1 V during rotation. The tachometer signal was also recorded by the three mentioned acquisition devices.

2.5 Test Plan.

To evaluate the influence of tip clearance, it is necessary to analyze different clearances at the same operating conditions and the same clearance at different operating conditions. This can be achieved due to the versatility of the facility. The turbine can be operated at a single rotational speed with different operating conditions as well as multiple rotational speeds for a given operating setpoint. For any combination of rotational speed and operating conditions, the high frequency instrumentation can be used to identify the impact of the blade-to-blade tip clearance variation on the local flow field. Additionally, tip clearances can be changed without modifying rotational speed or the operating condition through the thermal expansion of the blades and casing. These approaches help to isolate the effects of the thermal and mechanical stresses on the running tip clearance and provide insight into the impacts of the running clearance on the local flow field.

The turbine was operated at four total-to-static pressure ratios (labeled A, B, C, D) as shown Fig. 3. At each pressure ratio, the turbine was operated at four equivalent rotational speeds (labeled 80%, 90%, 100%, 110%), resulting in 16 total datasets. The equivalent rotational speed is defined as [36]
(1)
Fig. 3
Tests performed and conditions tests
Fig. 3
Tests performed and conditions tests
Close modal
Being
(2)
(3)
And Vcr,reference is evaluated at standard air temperature and pressure, with γreference=1.4. The equivalent speeds percentages result from their normalization by the nominal At-Design-Point of the turbine geometry tested
(4)

This array of test points allows the study of the same running clearance at different operating conditions, and the impact of the rotational speed on the running clearance at various operating conditions. In addition, run B included 10 min of dwell time at the nominal equivalent speed. This served as a temperature equilibrium test, where any observed change in tip clearance was attributed to thermal expansion of the blades and the shroud. This data addresses the scenario where the running tip clearance is changed while the turbine operating conditions and rotational speed are kept constant. It should be noted that the tip clearances can also be changed by swapping blade tracks, but this requires a new rig build and was outside the scope of the present study.

3 Data Reduction

3.1 Phase-Lock Averaging.

The high frequency data were phase-locked averaged (PLA) over the revolution period to retrieve the periodic physics. The PLA routine used was based on the instantaneous frequency estimation algorithm presented by Urbanek et al. [37]. This type of time-synchronous averaging was chosen over traditional low pass filtering to avoid the damping effects of small signal features and filtering artifacts as illustrated in Fig. 4. A preliminary recorded signal (top) was low pass filtered with a cutoff frequency ten times the blade passing frequency (BPD) (center). In the bottom figure, the data were phase-lock averaged, including more than 1500 revolution in the averaging. From this comparison, it is evident that the filtered results capture the general trend but fail to maintain the small details that characterize the flow physics.

Fig. 4
Comparison between raw pressure signal (top), low-pass filtered signal (center) and phase-locked averaged signal (bottom)
Fig. 4
Comparison between raw pressure signal (top), low-pass filtered signal (center) and phase-locked averaged signal (bottom)
Close modal
The PLA approach enables the identification of blade-to-blade variations and accounts for potential transients and oscillatory behaviors in the rotational speed due to variations in set speed or torque demanded by the dyno controller. The superimposition of all the periods included in the PLA calculation, as represented in Fig. 4, also enables the interrogation of the local data distribution and allows for the calculation of statistical quantities like the standard deviation throughout the rotation phase. In the example in Fig. 4, the 95% confident interval has been constructed using the calculated standard deviations. The standard error of the mean (SEM) is finally calculated as
(5)

Recording and including a minimum of 1000 revolution in the PLA reduces the standard error of the mean by two orders of magnitude. Table 1 provides the mean standard deviations and standard error on the means for the three measurements of interest. The largest values are observed in the heat flux signals, although they are still two orders of magnitude lower than the averaged signal. Therefore, the remaining data point plots of this paper do not include visible error bands.

Table 1

Standard deviation and standard error on the mean for the different measured magnitudes

Measurementσσ/Nrevs(>1000)
Tip clearance2.8%<0.09%
Shroud static pressure0.53%<0.016%
Shroud convective heat flux10.8%<0.34%
Measurementσσ/Nrevs(>1000)
Tip clearance2.8%<0.09%
Shroud static pressure0.53%<0.016%
Shroud convective heat flux10.8%<0.34%

3.2 Blade Identification and Signal Alignment.

Blade one was defined as the first entire blade at TDC after the tachometer trigger signal. The remaining blades are identified sequentially as they pass through TDC, that is, opposite to the spin direction.

An example of the resulting alignment is displayed Fig. 5(a). From top to bottom, this figure provides the signals of the voltage reading from the tip clearance measurements, the unsteady static pressure and heat flux. The magnitudes have been nondimensionalized as a fraction of the min and max measurements in the data window. The 95% confident interval constructed as two standard deviations is also provided.

Fig. 5
(a) Blade identification and signal alignment and (b) augmented view of three blades to identify key features (bottom)
Fig. 5
(a) Blade identification and signal alignment and (b) augmented view of three blades to identify key features (bottom)
Close modal

Figure 5(b) provides an augmented view of the aligned signals for three blade phases. The location of the suction and pressure side rims are clearly identified from the local maximum featured by the capacitance probe recording. This allows the identification of characteristic values associated with these locations and retrieval of the blade row signature.

3.3 Row Signatures Construction.

As presented in the top plot of Fig. 6, the slight blade-to-blade clearance variations results in a characteristic profile in the clearance for the entire row of blades, identified in this text as row signature. To analyze the clearance influence in the measured aerothermal properties, additional row signatures are constructed with characteristic magnitudes observed from the mutually aligned signals. Identified in Fig. 5(b), these are the magnitudes at the pressure and suction side rims, as well as the relevant feature found in the flow passage: a local maximum in the static pressure and a local minimum in the heat flux. These last ones are used to analyze whether the tip clearance has any influence on the midpassage flow in addition to the over tip flows. Figure 6 also exemplifies the row signatures for the pressure side rim at several equivalent rotational speeds.

Fig. 6
Blade row signatures at different equivalent speeds for the tip clearance voltage signal (top), shroud static pressure (middle) and surface heat flux (bottom)
Fig. 6
Blade row signatures at different equivalent speeds for the tip clearance voltage signal (top), shroud static pressure (middle) and surface heat flux (bottom)
Close modal

The magnitudes in this plot are expressed as a differential value to show the variation with the increasing operational speed. It is observed that the blade rows provide a characteristic signature profile. This signature is generally consistent for every condition, allowing a robust analysis of the features of interest.

4 Results and Discussion

Figures 5 and 6 provide a preliminary understanding of the overtip flows. The former provides time-resolved results. Starting from the main passage toward the pressure side of the blade, the shroud’s static pressure gradually decreases, identifying a flow acceleration toward the blade tip. The progressively increasing flow speed also results in an enhancement in convective heat flux with the shroud, with a peak value localized at the bladés pressure side rim, where the static pressure is minimum. This agrees with the simulations and observations performed by Ameri et al. [38], Lee et al. [39] and Bunker et al. [40] The flow then undergoes an expansion and deceleration after the first rim, reflected in a decrease in convective heat transfer. Within the tip cavity, several interactions of the leakage flow and pre-existing flow structures lead to some oscillations in both studied magnitudes as reported by Zou et al. [41], Maesschalk et al. [42], and Mischo et al. [4]. The flow eventually exits the tip region feeding into the tip leakage vortex. At this location, the static pressure gradually increases toward the main passage. Oppositely, the heat flux experiences a sudden reduction prior to the gradual increase toward the following blade tip. This sudden local drop in heat flux hints at an abrupt change of adiabatic wall temperature as previously observed by Lavagnoli et al. [22].

Figure 6 reveals that the increasing rotational speeds results in tighter tip clearances due to the larger mechanical loads on the blades. In this transition, the static pressure at the pressure side rim doesn’t show the same trend, as it increases when speeding up to 90% equivalent speed and decreases afterwards. The heat flux also features a different pattern than the tip clearance, but it opposes the behavior of the heat flux: it shows a decrease for the first speed transition and then it increases for the remaining two conditions.

A deeper analysis of all the magnitudes is performed in Secs. 4.14.5, identifying the potential governing relationships.

4.1 Tip Clearance.

To evaluate a potential dependance with tip clearance, it is necessary to analyze first the variation of this parameter at the various performance points. This is provided in Fig. 7. The left plot shows the variation in tip clearance referenced to the minimum value measured. Each dataset contains all the blades clearances at a 130 deg ALF circumferential location measured at the respective operating equivalent speed ratio.

Fig. 7
Evolution of tip clearance for the different tests
Fig. 7
Evolution of tip clearance for the different tests
Close modal

The increasing mechanical loads result in a difference close to 100 microns between the slowest and the overspeed conditions. The difference in clearance between runs results from the operating history. Runs A and B were performed consecutively in the same test; therefore, the undergoing thermal expansion of blades during the first run is noticeable in the slightly lower clearances featured in run B. The right plot provides the evolution of the clearance during the temperature equilibrium test after run B. It is apparent that the thermal expansion of the casing widens the clearance size over time, reaching a total increase of 10 microns. Run C and D were also performed within the same tests. In this case, the larger pressure ratios require higher operating temperatures to match the equivalent speed. The earlier thermal strains are observed from the tighter clearances than run A and B. Additionally, the shroud thermal expansion is also noticeable in run D.

4.2 Time-Resolved Unsteady Pressure.

The time-resolved shroud's unsteady pressure at the different operating conditions is provided in Fig. 8. Each plot corresponds to the different operating pressure ratios, and the magnitudes are normalized with the shroud's static pressure at the entrance of the rotor row (between stator and rotor) Two blade phases are illustrated, with the various equivalent speeds in different colors.

Fig. 8
Evolution of the shroud static pressure signal for all the tests
Fig. 8
Evolution of the shroud static pressure signal for all the tests
Close modal

For all the conditions, increasing rotational speeds results in larger flow speeds and consequently a decrease in static pressure. Oppositely, the increase of pressure ratio within the tip region for increasing rotational speed indicates a potential dependance with tip clearance that needs to be further investigated.

The bottom plot corresponds to the temperature equilibrium dwell time. A slight decrease in magnitude is observed from the local maximum in the main passage to the suction side tip rim of the previous blade. This also suggests a dependance with the variation in clearance.

An additional unsteadiness is revealed as a series of small oscillations throughout the entire signal for all conditions tested. These are more evident within the tip region and the pressure side region of the main passage. No reference to these trends was found in the previous studies, and without explicit reaction to the different operating conditions, no influence of tip clearance is inferred in this behavior. However, being that these profiles are a result from the phase-lock averaging over more than 1000 revolutions, it is suspected that these oscillations represent a real physical phenomenon related to the tight running clearances and high rotational speeds of the small-core turbine.

4.3 Time-Resolved Unsteady Heat Flux.

The homologous analysis performed for the convective surface heat flux measurements is provided in Fig. 9. At the pressure side rim, increasing the equivalent speed results in local heat flux enhancement. This is common for all the runs except for run A, where the trend is the opposite. Considering the lower operating pressure ratio of this condition, the resulting decrease of heat flux is explained by a reduction of leakage mass flow as the clearance tightens. The same enhancement behavior is noticed in the suction side rim. However, it is observed that the values at 110% equivalent speed are lower than at nominal speed, which can also be explained by a reduction of over-tip mass flow at the overspeed condition.

Fig. 9
Evolution of the shroud surface heat flux for all the tests
Fig. 9
Evolution of the shroud surface heat flux for all the tests
Close modal

The tip cavity profile shows an evolution with the turbine pressure ratio operating condition. Run A presents one local maximum in this region regardless of the equivalent speed. This extremum shifts toward the suction side rim in run B, and is no longer present in run C. The mismatch between these profiles and the previous surface pressure signals indicates that the mechanisms that govern the heat transfer are dependent on the operating pressure ratio, whereas the heat flux magnitude itself is influenced by clearance size.

The local minimum in the suction side is observed to differ between runs. For all tests, the location shifts toward the suction side with increasing speed. However, run A shows a larger magnitude with increasing speed, run B is steady, and runs C and D decrease. These trends suggest the existence of some compressibility effect being performed by the moving blade on the forward flow, as extensively discussed by Thorp et al. [43].

The temperature equilibrium case displays a general positive shift of the heat flux. Two effects contribute to this phenomenon. First, upstream temperature equilibrium reduces the upstream heat losses, therefore the flow has more energy at the measuring location and the heat flux therefore increases. Second, tip clearance widens, allowing more leakage flow and an enhancement in heat flux.

A better understanding of the governing physics behind the variations of heat flux is obtained by retrieving the local adiabatic wall temperature, adiabatic heat transfer coefficient, and Nusselt number. While the heat transfer mechanisms are governed by the recovery temperature [44,45], the adiabatic wall temperature provides a qualitative analysis of the former as they can be related through the Mach number [43]. The procedure to retrieve the first two has been described by Cernat and Lavagnoli [22,46] and is illustrated in the top two plots of Fig. 10. For every phase location, the evolution of the local heat flux at different wall temperatures is tracked, defining the linear relationship
(6)
Fig. 10
(a) Local phase measurements over time, (b) linear regression to retrieve adiabatic wall temperature and coefficient, and (c) resulting adiabatic wall temperature, convective coefficient and Nusselt number
Fig. 10
(a) Local phase measurements over time, (b) linear regression to retrieve adiabatic wall temperature and coefficient, and (c) resulting adiabatic wall temperature, convective coefficient and Nusselt number
Close modal

Therefore, the adiabatic convective heat transfer coefficient is found as the slope of the linear fit, and the adiabatic wall temperature is retrieved at null heat flux.

The temperature equilibrium study at the end of run B was used to determine the adiabatic wall properties as constant operating conditions were maintained and the clearance increase is minimum as presented earlier in Fig. 7. Therefore, the flow properties can be assumed constant in this period and the only variable is the shroud temperature.

The resulting profiles for the adiabatic wall temperature and heat transfer coefficient are displayed in Fig. 10(c). The former has been normalized with respect to the turbine total inlet temperature and the latter by its own mean value. This figure also contains the achieved Nusselt profile, computed as
(7)

In this formulation, kair is the thermal conductivity of air, and the local characteristic length was chosen to be the local hydraulic diameter Dh(φ), whose values are in very close proximity to the local clearance. With this selection, the characteristic length drastically changes between the flow passage, the blade tip rims and blade tip cavity. The large changes in this magnitude result in local Nusselt values of different orders of magnitude, as shown by the logarithmic scale on the y-axis of the corresponding plot in Fig. 10(c).

The adiabatic wall temperature (the driving temperature) features a sudden drop right before the suction side of the incoming blade, matching the expansion and recirculation of the tip leakage vortex. The amount of expansion in this location results in driving temperature values below the shroud wall temperature, reversing the heat flux direction and producing a cooling work from the airflow to the shroud. Toward midpassage the driving temperature rises gradually above the wall temperature and heat is transferred back toward the shroud. The enhancement in heat transfer in the tip region follows the increase of the driving temperature in this area. The adiabatic wall temperature doesn't display values above the turbine inlet total temperature, opposing elevated compressive work levels reported in Refs. [22,43]. This disagreement suggests that the passage of small-core small aspect ratio turbines is highly influenced by the developed detached secondary flows.

The adiabatic heat transfer coefficient also features localized larger values within the tip region, indicating a different flow field than the main flow passage. Local maximums are observed matching the tip rims, the locations of tightest clearance and peak flow speed. The oscillations of this coefficient within the cavity correspond to the different flow expansions and recirculation regions.

The drastic changes in Nusselt number support the concept of different flow fields as the blades pass. Within the main passage, the local Nusselt was compared with the solution proposed by Bhatti and Shah [47] for turbulent flow in ducts
(8)

Excellent match was found by assuming a fully developed flow around 90% of the blade chord, agreeing with the results provided by Solano et al. [48]. Within the tip region, no correlation in the open literature was found to provide a good match with the obtained results, although the values agree with the tip measurements of some previous studies [49]. This indicates the presence of highly complex flow structures within squealer tips.

To further relate the heat flux variations with the changes in fluid temperature or disturbances in the boundary layer, the measured heat flux is expressed as the addition of steady and unsteady components as follows:
(9)
The unsteady terms are then clearly identified as a contribution from an adiabatic temperature change (qT), a contribution from an adiabatic convective heat flux coefficient change (qh), this one denoting a variation in boundary layer [50], and a third term related to the unsteadiness (qh&T)
(10)
(11)
(12)
(13)

The profiles of the different contributions are displayed in Fig. 11 and their influence into the overall heat flux and overall unsteady heat flux is indicated in Table 2. The magnitude of the unsteady components contributes to almost half of the overall heat flux. This is seen in the normalized profile shown in Fig. 11, depicting the unsteady component to periodically oscillate between −150% and 150% of the steady value. Within the unsteady heat flux, the changes in the adiabatic wall temperature display the largest influence, accounting for one-third of the overall heat flux processes. Variations in the boundary layer represent almost 20% of the overall unsteady heat flux. The lowest contribution is displayed by the qh&T term, also included in the lowest plot of Fig. 11, with an influence of only 3.5% of the total heat transfer processes.

Fig. 11
Unsteady heat flux components
Fig. 11
Unsteady heat flux components
Close modal
Table 2

Contribution of the unsteady heat flux components

Q˙=q¯+q=q¯+qh+qT+qh&TInfluence on Q˙Influence on q
q¯55.75
q44.25
qh+haw(Taw¯Twall)6.25%18.3%
qT=haw¯Taw34.5%74.1%
qh&T=hawTaw3.5%7.6%
Q˙=q¯+q=q¯+qh+qT+qh&TInfluence on Q˙Influence on q
q¯55.75
q44.25
qh+haw(Taw¯Twall)6.25%18.3%
qT=haw¯Taw34.5%74.1%
qh&T=hawTaw3.5%7.6%

Note: The bold values represent the overall steady and unsteady components.

4.4 Comparison With Computational Heat Flux Prediction.

To evaluate the limitations of commonly used computational tools, RANS simulations were performed for the turbine geometry and compared against the empirical measurements. Computational fluid dynamics (CFD) simulations of the entire 2 stage turbine were performed using ANSYS Fluent 2023R2. The experimentally obtained tip gap height was used in the CFD model. The mesh, represented in Fig. 12(a), was generated using the embedded meshing tool in Fluent. An unstructured polyhedral mesh in the core and prism layers at the wall was generated securing a y+ value strictly below 1. A mesh convergence study was executed to identify a sufficient resolution of the core mesh. The tip mesh has at least 44 cells in radial direction at any location of the tip. The k − ω shear stress transport turbulence model was applied. The boundary conditions for inlet turbulence intensity, total inlet temperature, and total inlet pressure, as well as the static outlet pressure were set in accordance with the experimental values for the run B dwell time.

Fig. 12
(a) Blade tip computational mesh and (b) superimposition of experimental measurements and computational convective surface heat flux
Fig. 12
(a) Blade tip computational mesh and (b) superimposition of experimental measurements and computational convective surface heat flux
Close modal

Figure 12(b) displays the computational and experimental results for the shroud surface convective heat flux normalized by the respective means. The location of the blade tip in the predictions was returned by the computational tool and aligned with the experimental measurements. Two sets of computational results are provided: one at the axial chord location matching the experimental measurements and a second set resulting from its averaging over the ALTP sensor measuring area. There is good agreement between the results in the main passage region. A heat flux reduction is experienced near the suction side and progressively increases toward the pressure side. The predicted peak heat flux is located at the entrance of the pressure side rim, coinciding with the maximum pressure drop rate. Although the measured peak heat flux is lower than the computational punctual result, very good agreement is observed when compared against the area averaged prediction. However, the experimental maximum is found further into the blade rim, suggesting the existence of a vena contracta as reported by Moore et al. [8], Lee et al. [39], and Ameri et al. [2]. Within the blade cavity, the simulations infer a larger flow deceleration and heat flux reduction prior to the suction side rim, where a local maximum is again predicted. This trend is observed also in the area-averaged results at a lower rate. The experimental results indicate a lower flow deceleration and a more progressive reduction toward the suction side rim. This mismatch suggests a limitation of the computational tool to precisely resolve the complex detached flows within the cavity region and evidences the necessity for the complimentary experimental data.

4.5 Effect of Tip Clearance in Static Pressure.

The blade row signatures provide a first analysis on a potential tip clearance influence. This is illustrated in the top three plots of Fig. 13, providing the normalized pressure values at the main passage, suction, and pressure side rim. As aforementioned, the pressure in the main passage decreases with increasing rotational speed, whereas the values at the blade tip rims increase. The conservation of the signature profile at every location and condition indicates that the tip clearance influences the pressure field both within the blade tip and the main passage regions.

Fig. 13
Top: pressure row signatures of the different signal relevant features; bottom: signature of the pressure ratio across the tip for runs A and D
Fig. 13
Top: pressure row signatures of the different signal relevant features; bottom: signature of the pressure ratio across the tip for runs A and D
Close modal
To characterize the leakage profile through each blade, the following pressure ratio was evaluated for each blade:
(14)

Negative values indicate lower pressure at the pressure side rim than the suction side rim, indicating flow acceleration from the main passage toward the pressure side rim and over-tip flows from the pressure side toward the suction side. Oppositely, positive values result from pressures in the pressure side rim above the magnitude on the suction side rim, and therefore higher flow speeds in the latter location. This scenario indicates the existence of reverse over-tip flows from the suction side toward the pressure side. The bottom plots of Fig. 13, exemplify these row signatures for the runs of lowest and highest operating turbine pressure ratio (runs A and D). In run A, only the slowest equivalent speed features negative pressure ratios. As the equivalent speed increases and the clearance decreases, the pressure ratios become positive and barely change with increasing speed. In the case of run D, lower values are experienced at the slowest speed, increasing with higher spin, and positive pressure ratios are only displayed at the largest equivalent speed and at some blades at nominal speed. Therefore, it is inferred that for a given operating turbine pressure ratio, there exists a clearance threshold below which the over-tip flow structures change, and reverse over-tip flows are experienced as predicted by the simulations of Andreoli [51] and de-Maesschalck [11].

To analyze this threshold, the tip pressure ratio was analyzed against the performance condition of the blade tip itself, for which the following rotational tip Reynolds number was defined:
(15)

For this Reynolds definition, the tip clearance (C) was considered a better choice for the characteristic length than the blade tip radius as it is more realistic representation of the flow within the tip clearance. As portrayed in Fig. 14 (left), well-defined clusters for each equivalent speed linearly increase with tighter tip clearance. A linear fit through each group allows computing the transitional Reynolds. This threshold is represented against the equivalent speed in Fig. 14 (right). The revealed linear dependance between these two parameters provides a solid reference for turbine design, providing knowledge on the leakage flow structure from the rotational tip Reynolds.

Fig. 14
Left: pressure ratio across blade tip against rotational tip Reynolds; right: critical rotational tip Reynolds against equivalent rotational speed
Fig. 14
Left: pressure ratio across blade tip against rotational tip Reynolds; right: critical rotational tip Reynolds against equivalent rotational speed
Close modal

4.6 Effect of Tip Clearance in Surface Heat Flux.

The same analysis approach was used to investigate heat flux dependances on tip clearances. The top three plots of Fig. 15 illustrate the heat flux signatures referenced to the absolute minimum measured for the key local features. While these plots seem to maintain comparable signatures at different speeds, it isn't as remarkable as the previous pressure signatures, indicating only a partial influence of tip clearance on the different magnitudes. An additional relevant outcome from these plots is the position of the outlaying feature. This one corresponds to blade 4 on the pressure side and suction side and blade 3 in the passage signature (recall that the passage feature was grouped with the forward blade). This shift indicates that the magnitude of the sudden decrease in heat flux at the suction side tip is dependent on the clearance of such blade tip and not on the forward blade clearance like the pressure profile.

Fig. 15
Heat flux row signatures of the different signal relevant features; bottom: evaluation against the rotational tip Reynolds for runs A and D
Fig. 15
Heat flux row signatures of the different signal relevant features; bottom: evaluation against the rotational tip Reynolds for runs A and D
Close modal

The peak heat flux in the blade tip is then evaluated against the rotational tip Reynolds. Figure 15 illustrates again the lowest and highest-pressure ratio runs (A and D). Like the pressure analysis, each equivalent speed displays a defined bundle. However, increasing speed doesn't result in a common shift for both sets of data. Within each speed ensemble, a positive linear relationship is discerned with the tip rotational Reynolds. The low slopes of these trends indicate an existing, but low contribution of tip clearance in the determination of the heat flux magnitude.

A deeper knowledge of tip clearance influence is obtained by analyzing its relationship with the different unsteady heat flux components as portrayed in Fig. 16. Figure (a) identifies the characteristic features being studied for each contribution (pressure and suction side rims, and the location of local minimum of heat flux prior to the suction side). Figure (b) provides the evaluation against the tip clearance. This figure only contains data for the dwell time postrun B as only this one allowed retrieving the unsteady heat flux components. The results reveal linear trends between the overall unsteady heat flux and the tip clearance value for the tip region (pressure and suction side rims). It is observed that the heat flux due to boundary layer variations and the unsteady term hawTAw show constant values regardless of the clearance size. The tip clearance shows the largest influence on the adiabatic wall temperature unsteady heat flux. These results show a reduction of unsteady heat flux in the pressure side rim with decreasing clearance, and an opposing trend in the suction side rim. The values of heat flux upstream to the suction side show no correlation with the tip clearance, resulting in a homogeneous range for the entire span of clearances. These behaviors again agree with the transition to reverse over-tip flows and reductions in leakage flow with smaller clearances predicted in Refs. [11] and [51].

Fig. 16
(a) Features to study for each unsteady heat flux component and (b) characteristic features against tip clearance
Fig. 16
(a) Features to study for each unsteady heat flux component and (b) characteristic features against tip clearance
Close modal

4.7 Coupling Between Pressure and Heat Flux.

A final study is performed to investigate potential relationships between the measured static pressure fields and convective heat flux. To that end, a first analysis was performed between the unsteady heat flux components and the blade tip pressure ratio as computed per Eq. (14). The results are displayed below in Fig. 17, indicating again in the left the signal features being evaluated. The linear trends observed in the overall unsteady heat flux indicate an increase in heat load in the suction side rim and opposing decay in the pressure side rim for increasing pressure ratio. Again, this trend supports the existence of reverse flow structures as more positive pressure ratios result in larger overtip flow speeds in the suction side compared to the pressure side. No additional dependance is noticeable between the local minimum prior the suction side with tip pressure ratio.

Fig. 17
(a) Features to study for each unsteady heat flux component and (b) evaluation against static pressure
Fig. 17
(a) Features to study for each unsteady heat flux component and (b) evaluation against static pressure
Close modal

A second analysis was performed to investigate a potential governance between the static pressure and the adiabatic heat transfer coefficient contribution, as small oscillations where observed in both (visible in Fig. 8 for the static pressure and the bottom plot of Fig. 11 for the heat flux component). To the best of the author's knowledge, this phenomenon has not been reported in the open literature. A frequency analysis was performed to assess the potential correlation between both effects. As the PLA only returns features that are periodic to the averaging period, the presence of these oscillations indicates a phenomenon that is a harmonic of the blade passing frequency. Figure 18(a) provides superimposition of the pressure signal and the unsteady component qh. Figure 18(b) presents the frequency components for both signals, where the x-axis has been normalized by the blade passing frequency to return the harmonic numbers, and the y-axis is expressed in decibels. Aside from the BPD, the frequency content of the heat flux contribution is always larger than the harmonic contributions of the pressure signal. It is noticed that the harmonics 9 and 10 display very similar contribution for both signals, and these numbers match the number of oscillations counted within the blade periods displayed in Fig. 18(a). However, this match does not display any predominant weight with respect to the remaining harmonics. Consequently, it is not possible to assure a robust correlation between the high-frequency low-magnitude components observed in the static pressure and the studied heat flux component.

Fig. 18
(a) Normalized profiles and (b) frequency contents of the static pressure and convective coefficient unsteady heat flux
Fig. 18
(a) Normalized profiles and (b) frequency contents of the static pressure and convective coefficient unsteady heat flux
Close modal

5 Conclusions

This paper has presented a comprehensive experimental assessment of the shroud's unsteady static pressure and heat flux for a small core squealer-tip-blades turbine. A suitable facility and test section allowed performing the empirical measurements at engine representative conditions of pressure ratio, equivalent speeds, and tip clearance values. The proper phase-lock averaging of all the measurements allowed retrieving the true periodic physics behind each magnitude with standard errors of the mean below 0.35% of the measured value.

The evaluation against the computational simulations has evidenced the necessity of complimentary experimental data to validate the performance of unsteady simulations in resolving highly detached flows in low aspect ratio turbomachinery.

The obtained results have revealed a large influence of the tip clearance size in the over-tip flow characteristics. A direct relationship has been observed between the rotational tip Reynolds and the pressure ratio through the tip clearance region. It has been evidenced that for a given equivalent rotational speed, there is a critical tip Reynolds, and therefore clearance, below which reverse over-tip flows are experienced.

The analysis of the shroud's unsteady heat flux revealed an outstanding contribution of the unsteady terms into the overall over-tip shroud heat flux, amounting to almost 45% of contribution. Within the unsteady heat flux components, a large dominance was noticed from the variations in the adiabatic wall temperature, resulting in 1/3 of the total influence on the overall heat flux. Posterior analysis revealed that tip clearance size has a direct influence on the enhancement of this specific unsteady heat flux component within the tip region.

Final inspections between the shroud's static pressure and heat flux supported the existence of reversed over-tip flows resulting in heat load enhancement in the suction side rim and reduction in the pressure side rim.

Acknowledgment

All the empirical work presented in this text was assisted by the members of the STARR team within the PETAL research group as well as the project sponsor Rolls Royce members, Matthew Bloxham and Eugene Clemens, who provided guidance and insightful discussion on the several findings.

Data Availability Statement

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

Nomenclature

C =

tip clearance

Dh =

hydraulic diameter

haw =

adiabatic convective heat transfer coefficient

N =

rotational speed

Nrevs =

number of revolutions

Nsamples =

number of samples

Nu =

Nusselt number

Nuz =

circumferential local Nusselt number

Nu =

fully developed flow Nusselt number

N¯ =

equivalent rotational speed

P =

static pressure measured signal

PPASS =

maximum static pressure in the flow passage

PPS =

static pressure at the pressure side rim

PSS =

static pressure at the suction side rim

P2,Stg1 =

static pressure between vane and blade in Stg1

PRTS =

total-to-static turbine pressure ratio

Q˙ =

heat flux

R =

ideal gas constant

Rshroud =

shroud radius

Rtip =

tip radius

ReTIP =

rotational Reynolds number at the blade tip

Taw =

adiabatic wall temperature

Twall =

wall temperature

T0 =

total temperature

VC =

tip clearance voltage signal

x =

axial location

γ =

specific heat ratio

ΠTIP =

static pressure ratio across the blade tip

σ =

standard deviation

φ =

phase

Ω¯ =

equivalent rotational speed ratio

Acronyms
ALF =

aft looking front

CFD =

computational fluid dynamics

PLA =

phase-locked average

RANS =

Reynolds averaged Navier–Stokes

SEM =

standard error on the mean

TDC =

top dead center

References

1.
Metzger
,
D. E.
,
Dunn
,
M. G.
, and
Hah
,
C.
,
1991
, “
Turbine Tip and Shroud Heat Transfer
,”
ASME J. Turbomach.
,
113
(
3
), pp.
502
507
.10.1115/1.2927902
2.
Ameri
,
A. A.
,
2001
, “
Heat Transfer and Flow on the Blade Tip of a Gas Turbine Equipped With a Mean-Camberline Strip
,”
ASME J. Turbomach.
,
123
(
4
), pp.
704
708
.10.1115/1.1400114
3.
Denton
,
J. D.
,
1993
, “
Loss Mechanism in Turbomachines
,”
ASME
Paper No. 93-GT-435.10.1115/93-GT-435
4.
Mischo
,
B.
,
Behr
,
T.
, and
Abhari
,
R. S.
,
2008
, “
Flow Physics and Profiling of Recessed Blade Tips: Impact on Performance and Heat Load
,”
ASME J. Turbomach.
, 130(2), p.
021008
.10.1115/1.2775485
5.
Li
,
W.
,
Jiang
,
H.
,
Zhang
,
Q.
, and
Woo Lee
,
S.
,
2014
, “
Squealer Tip Leakage Flow Characteristics in Transonic Condition
,”
ASME J. Eng. Gas Turbine Power
,
136
(
4
), p.
042601
.10.1115/1.4025918
6.
Kegalj
,
M.
,
Schmid
,
G.
,
Wartzek
,
F.
,
Schiffer
,
H.-P.
, and
Willer
,
L.
,
2012
, “
Experimental and Numerical Investigation of Tip Leakage Flow in a 1 1/2 Stage Turbine Rig Comparing Flat and Cavity-Squealer Tip Geometries
,”
ASME
Paper No. GT2012-69568.10.1115/GT2012-69568
7.
Meng
,
F.
,
Zheng
,
Q.
,
Gao
,
J.
, and
Fu
,
W.
,
2019
, “
Effect of Tip Clearance on Flow Field and Heat Transfer Characteristics in a Large Meridional Expansion Turbine
,”
Energies (Basel)
,
12
(
1
), p.
162
.10.3390/en12010162
8.
Moore
,
J.
, and
Tilton
,
J. S.
,
1988
, “
Tip Leakage Flow in a Linear Turbine Cascade
,”
ASME J. Turbomach.
,
110
(
1
), pp.
18
26
.10.1115/1.3262162
9.
Moore
,
J.
, and
Elward
,
K. M.
,
1993
, “
Shock Formation in Overexpanded Tip Leakage Flow
,”
ASME J. Turbomach.
,
115
(
3
), pp.
392
399
.10.1115/1.2929266
10.
Papa
,
M.
,
Goldstein
,
R. J.
, and
Gori
,
F.
,
2003
, “
Effects of Tip Geometry and Tip Clearance on the Mass/Heat Transfer From a Large-Scale Gas Turbine Blade
,”
ASME J. Turbomach.
,
125
(
1
), pp.
90
96
.10.1115/1.1529190
11.
De Maesschalck
,
C.
,
Lavagnoli
,
S.
,
Paniagua
,
G.
, and
Vinha
,
N.
,
2014
, “
Aerothermodynamics of Tight Rotor Tip Clearance Flows in High-Speed Unshrouded Turbines
,”
Appl. Therm. Eng.
,
65
(
1–2
), pp.
343
351
.10.1016/j.applthermaleng.2014.01.015
12.
Teia
,
L.
,
2019
, “
New Insight Into Aspect Ratio's Effect on Secondary Losses of Turbine Blades
,”
ASME J. Turbomach.
,
141
(
11
), p.
111004
.10.1115/1.4044079
13.
Rogo
,
C.
,
1968
, “
Experimental Aspect Ratio and Tip Clearance Investigation on Small Turbines
,”
SAE Transactions
, 77, pp.
1726
1743
.http://www.jstor.org/stable/44565245
14.
Andreoli
,
V.
,
Paniagua
,
G.
, and
Bloxham
,
M.
,
2021
, “
Towards Desensitization of Gas Turbine Performance to Tip Clearance: Design Optimization and Engine Analysis
,”
Energy Convers. Manage.
,
245
, p.
114575
.10.1016/j.enconman.2021.114575
15.
Zhong
,
B.
,
Scheurich
,
F.
,
Titarev
,
V.
, and
Drikakis
,
D.
,
2009
, “
Turbulent Flow Simulations Around a Multi-Element Airfoil Using URANS, DES and ILES
Approaches,”
AIAA
Paper No. 2009-3799.10.2514/6.2009-3799
16.
Saavedra
,
J.
, and
Paniagua
,
G.
,
2019
, “
Transient Performance of Separated Flows: Experimental Characterization of Flow Detachment Dynamics
,”
ASME
Paper No. GT2019-91020.10.1115/GT2019-91020
17.
Thorpe
,
S. J.
,
Yoshino
,
S.
,
Ainsworth
,
R. W.
, and
Harvey
,
N. W.
,
2004
, “
An Investigation of the Heat Transfer and Static Pressure on the Over-Tip Casing Wall of an Axial Turbine Operating at Engine Representative Flow Conditions. (II). Time-Resolved Results
,”
Int. J. Heat Fluid Flow
,
25
(
6
), pp.
945
960
.10.1016/j.ijheatfluidflow.2004.02.028
18.
Wheeler
,
A. P. S.
,
Atkins
,
N. R.
, and
He
,
L.
,
2011
, “
Turbine Blade Tip Heat Transfer in Low Speed and High Speed Flows
,”
ASME J. Turbomach.
,
133
(
4
), p.
041025
.10.1115/1.4002424
19.
Zhang
,
Q.
,
O'Dowd
,
D. O.
,
He
,
L.
,
Wheeler
,
A. P. S.
,
Ligrani
,
P. M.
, and
Cheong
,
B. C. Y.
,
2011
, “
Overtip Shock Wave Structure and Its Impact on Turbine Blade Tip Heat Transfer
,”
ASME J. Turbomach.
,
133
(
4
), p.
041001
.10.1115/1.4002949
20.
Krishnababu
,
S. K.
,
Newton
,
P. J.
,
Dawes
,
W. N.
,
Lock
,
G. D.
,
Hodson
,
H. P.
,
Hannis
,
J.
, and
Whitney
,
C.
,
2009
, “
Aerothermal Investigations of Tip Leakage Flow in Axial Flow Turbines—Part I: Effect of Tip Geometry and Tip Clearance Gap
,”
ASME J. Turbomach.
,
131
(
1
), p.
011006
.10.1115/1.2950068
21.
Krishnababu
,
S. K.
,
Dawes
,
W. N.
,
Hodson
,
H. P.
,
Lock
,
G. D.
,
Hannis
,
J.
, and
Whitney
,
C.
,
2009
, “
Aerothermal Investigations of Tip Leakage Flow in Axial Flow Turbines—Part II: Effect of Relative Casing Motion
,”
ASME J. Turbomach.
,
131
(
1
), p.
011007
.10.1115/1.2952378
22.
Lavagnoli
,
S.
,
Paniagua
,
G.
,
De Maesschalck
,
C.
, and
Yasa
,
T.
,
2013
, “
Analysis of the Unsteady Overtip Casing Heat Transfer in a High Speed Turbine
,”
ASME J. Turbomach.
,
135
(
3
), p.
031027
.10.1115/1.4007509
23.
Paniagua
,
G.
,
Cuadrado
,
D.
,
Saavedra
,
J.
,
Andreoli
,
V.
,
Meyer
,
T.
,
Solano
,
J. P.
,
Herrero
,
R.
,
Meyer
,
S.
, and
Lawrence
,
D.
,
2019
, “
Design of the Purdue Experimental Turbine Aerothermal Laboratory for Optical and Surface Aerothermal Measurements
,”
ASME J. Eng. Gas Turbine Power
,
141
(
1
), p.
012601
.10.1115/1.4040683
24.
Cuadrado
,
D. G.
,
Aye-Addo
,
P. A. N.
,
Andreoli
,
V.
,
Bhatnagar
,
L.
,
Lozano
,
F.
,
Fisher
,
J.
,
Paniagua
,
G.
, et al.,
2019
, “
Purdue Small Turbine Aerothermal Rotating Rig (STARR)
,”
AIAA
Paper No. 2019-4004.10.2514/6.2019-4004
25.
Fogale Nanotech
, 2023, “
Capablade Fusion General Documentation: Capacitive Measurement System Dedicated to Blade Tip Clearance and Tip Timing Measurement
,” Fogale Nanotech, Nimes, France, accessed Dec. 1, 2023, www.fogale.fr
26.
Fogale Nanotech
, 2023, “
Technical Documentation MC925 Compatible With Capablade System Capacitive Measurement System Dedicated to Blade Tip Clearance and Tip Timing Measurement
,” Nimes, France, 2023, accessed Dec. 1, 2023, www.fogale.fr
27.
Salinas
,
S.
,
Castillo
,
A.
,
Paniagua
,
G.
, and
Bloxham
,
M.
,
2023
, “
Measurement Comprehensive Capacitance Sensor Calibration for High-Speed Fluid-Machinery Tip Clearance Characterization
,”
Measurements
,
218
, p.
113117
.10.1016/j.measurement.2023.113117
28.
Castillo Sauca
,
A.
, and
Paniagua
,
G.
,
2023
, “
Robust Tip Gap Measurements: A Universal In-Situ Dynamic Calibration & Demonstration In A Two-Stage High-Speed Turbine
,”
ASME J. Eng. Gas Turbine Power
,
146
(
6
), pp.
1
36
.10.1115/1.4063886
29.
XCE-062 Rev. P (3)
,”.
30.
Precision Filters Inc
,
2017
, “
Precision Model 101777: Triple Output Bridge Conditioner
,” Precision Filters Inc, Ithaca, NY, accessed Dec. 1, 2023, www.pfinc.com
31.
HBK
, 2024, “
B04649_04_E00_00 GEN Series GEN2tB Transient Recorder and Data Acquisition System
,” Hottinger Bruel & Kjaer UK Ltd, Norcross, GA.
32.
HBK
, 2024, “
B04596_03_00 GEN Series GN8101B/GN8102B/GN8103B Basic 250, 100, 25 MS/s Input Card
,” Hottinger Bruel & Kjaer UK Ltd, Norcross, GA.
33.
Wagner
,
S. N.
,
Krämer
,
E.
,
Roediger
,
T.
,
Knauss
,
H.
,
Wagner
,
S.
,
Kraemer
,
E.
,
Bountin
,
D. A.
, et al.,
2007
, “
The Atomic Layer Thermopile - A Fast Heat Flux Sensor for Measuring High Heat Loads in Short Duration Hypersonic Ground Testing Facilities
,” accessed Dec. 1, 2023, https://api.semanticscholar.org/CorpusID:220398372
34.
Pico Technology
,
2023, “
FlexRes® Oscilloscopes and MSOs PicoScope® 5000D Series
,” Pico Technology, St Neots, UK, accessed Dec. 1, 2023, www.picotech.com
35.
MCS-655M_Reflective_Sensor
,”.
36.
Bhatnagar
,
L.
,
Paniagua
,
G.
,
Clemens
,
E.
, and
Bloxham
,
M.
,
2024
, “
A Probabilistic Approach to Turbine Uncertainty
,”
ASME J. Turbomach.
,
146
(
4
), p.
041009
.10.1115/1.4064187
37.
Urbanek
,
J.
,
Barszcz
,
T.
, and
Antoni
,
J.
,
2013
, “
A Two-Step Procedure for Estimation of Instantaneous Rotational Speed With Large Fluctuations
,”
Mech. Syst. Signal Process.
,
38
(
1
), pp.
96
102
.10.1016/j.ymssp.2012.05.009
38.
Ameri
,
A. A.
,
Steinthorsson
,
E.
, and
Rigby
,
D. L.
,
1998
, “
Effect of Squealer Tip on Rotor Heat Transfer and Efficiency
,”
ASME J. Turbomach.
,
120
(
4
), pp.
753
759
.10.1115/1.2841786
39.
Lee
,
S. E.
, and
Lee
,
S. W.
,
2016
, “
Over-Tip Leakage Flow and Loss in a Turbine Cascade Equipped With Suction-Side Partial Squealers
,”
Int. J. Heat Fluid Flow
,
61
, pp.
575
584
.10.1016/j.ijheatfluidflow.2016.07.001
40.
Bunker
,
R. S.
,
Bailey
,
J. C.
, and
Ameri
,
A. A.
,
2000
, “
Heat Transfer and Flow on the First-Stage Blade Tip of a Power Generation Gas Turbine: Part 1—Experimental Results
,”
ASME J. Turbomach.
,
122
(
2
), pp.
263
271
.10.1115/1.555443
41.
Zou
,
Z.
,
Shao
,
F.
,
Li
,
Y.
,
Zhang
,
W.
, and
Berglund
,
A.
,
2017
, “
Dominant Flow Structure in the Squealer Tip Gap and Its Impact on Turbine Aerodynamic Performance
,”
Energy
,
138
, pp.
167
184
.10.1016/j.energy.2017.07.047
42.
De Maesschalck
,
C.
,
Andreoli
,
V.
,
Paniagua
,
G.
,
Gillen
,
T.
, and
Barker
,
B.
,
2021
, “
Aerothermal Optimization of Turbine Squealer Tip Geometries With Arbitrary Cooling Injection
,”
ASME J. Turbomach.
,
143
(
11
), p.
111010
.10.1115/1.4051268
43.
Thorpe
,
S. J.
,
Yoshino
,
S.
,
Ainsworth
,
R. W.
, and
Harvey
,
N. W.
,
2004
, “
An Investigation of the Heat Transfer and Static Pressure on the Over-Tip Casing Wall of an Axial Turbine Operating at Engine Representative Flow Conditions. (I). Time-Mean Results
,”
Int. J. Heat Fluid Flow
,
25
(
6
), pp.
933
944
.10.1016/j.ijheatfluidflow.2004.02.027
44.
Thorpe
,
S. J.
, and
Ainsworth
,
R. W.
,
2008
, “
The Effects of Blade Passing on the Heat Transfer Coefficient of the Overtip Casing in a Transonic Turbine Stage
,”
ASME J. Turbomach.
,
130
(
4
), p.
041009
.10.1115/1.2776950
45.
Thorpe
,
S. J.
,
Miller
,
R. J.
,
Yoshino
,
S.
,
Ainsworth
,
R. W.
, and
Harvey
,
N. W.
,
2007
, “
The Effect of Work Processes on the Casing Heat Transfer of a Transonic Turbine
,”
ASME J. Turbomach.
,
129
(
1
), pp.
84
91
.10.1115/1.2372772
46.
Cernat
,
B.
, and
Lavagnoli
,
S.
,
2024
, “
Experimental Investigation of the Tip Design Effects on the Unsteady Aerodynamics and Heat Transfer of a High Speed Turbine
,”
ASME J. Turbomach.
, 146(9), p.
091008
.10.1115/1.4050443
47.
Kakaç
,
S.
,
Shah
,
R. K.
, and
Aung
,
W.
,
1987
,
Handbook of Single-Phase Convective Heat Transfer
, John Wiley and Sons Inc, New York.
48.
Solano
,
J. P.
,
Pinilla
,
V.
,
Paniagua
,
G.
,
Lavagnoli
,
S.
, and
Yasa
,
T.
,
2011
, “
Aero-Thermal Investigation of a Multi-Splitter Axial Turbine
,”
Int. J. Heat Fluid Flow
,
32
(
5
), pp.
1036
1046
.10.1016/j.ijheatfluidflow.2011.05.011
49.
Chana
,
K. S.
, and
Jones
,
T. V.
,
2003
, “
An Investigation on Turbine Tip and Shroud Heat Transfer
,”
ASME J. Turbomach.
,
125
(
3
), pp.
513
520
.10.1115/1.1575253
50.
Pinilla
,
V.
,
Solano
,
J. P.
, and
Paniagua
,
G.
,
2011
, “
Method and Measurement of Adiabatic Wall Temperature Downstream of a High Pressure Turbine Using Thin-Film Sensors
,”
ASME
Paper No. GT2011-46556.10.1115/GT2011-46556
51.
Andreoli
,
V.
,
Cuadrado
,
D. G.
, and
Paniagua
,
G.
,
2018
, “
Prediction of the Turbine Tip Convective Heat Flux Using Discrete Green's Functions
,”
ASME J. Heat Mass Transfer-Trans. ASME
,
140
(
7
), p.
071703
.10.1115/1.4039182