Abstract

Next generation aeroengines will operate at ever-increasing pressure ratios with smaller cores, where the control of blade-tip clearances across the flight cycle is an emerging design challenge. Such clearances are affected by the thermal expansion of the compressor disks that hold the blades, where acute thermal stresses govern operating life. The cavities formed by corotating disks feature a heated shroud at high radius and cooler cobs at low radius. A three-dimensional, unsteady and unstable flow structure is induced by destabilizing buoyancy forces. The radial distribution of disk temperature is driven by a conjugate heat transfer at Grashof numbers of order 1013. Such flows are further influenced by the heat and mass exchange with an axial throughflow of cooling air at low radius, where the interaction depends on the Rossby number and separation of the disk cobs. This paper is the first to study the effect of cob separation ratio on mass and heat exchange for compressor cavities. A model is developed to predict the cavity-throughflow interaction, and disk and fluid-core temperatures. The judicious use of a physics-based methodology provides reliable, reduced-order solutions to the complex conjugate problem, thereby making it appropriate for practical engine thermo-mechanical design. The model is validated by detailed experimental measurements using the Bath Compressor Cavity Rig, where variable disk cob spacings were investigated over a range of engine-representative conditions. The unsteady pressure measurements collected in the frame of reference of the rotating disks reveal new insight into the fundamentally aperiodic nature of the flow structure. This new understanding of heat transfer informs an expedient reduced-order model and enables more efficient design of future high pressure-ratio aeroengines.

1 Introduction

Blade tip clearances and thermal stresses affect the efficiency and operating life of gas turbine compressors. Engine thermo-mechanical design and analysis requires accurate prediction of compressor disk temperatures, which are determined by the flow structure and heat transfer within the cavities formed by corotating disks and the rotor shroud. Figure 1 shows a typical aero-engine arrangement, with an axial throughflow of cool air in the bore annulus at low radius. In practice, most compressor cavities are open with thick cobs at the inner radii of the disks where hoop stresses are at a maximum [1]. The cavity shroud is heated by the mainstream annulus and temperature differences lead to buoyancy-induced flow. Such flows feature unsteady and three-dimensional vortex structures, ingress of cool fluid from the axial throughflow, enthalpy, and momentum exchange, and egress of hot fluid from the cavity.

Fig. 1
Typical cross section of a high-pressure aero-engine compressor showing key geometrical definitions
Fig. 1
Typical cross section of a high-pressure aero-engine compressor showing key geometrical definitions
Close modal
The cavity and cob dimensions are clearly important to the mass and heat interaction between the fluid in the rotating cavity and the axial throughflow. Some cavities are open to interaction between the cavity and the axial throughflow in the center, while others have cobs of varying thickness partially or fully preventing throughflow-cavity interaction. With reference to Fig. 1, a cavity can be characterized by five key geometrical parameters: cavity radius ratio (a/b), bore annulus radius ratio (rs/a), cob height ratio (a/b), cavity aspect ratio (G=s/b), and the cob separation ratio,
(1)
The throughflow-cavity flow interaction is a strong function of η and its effect on the heat transfer and flow structure in the compressor cavity is the focus of this paper. The two limiting cases are η=0 and η=1. When η=0, the cavity is closed and throughflow interaction is prevented. η=1 is a fully open cavity with maximum cavity-throughflow interaction. Due to the free convective, rotating, compressible nature of the flow, along with cavity-throughflow interaction, there are four governing dimensionless parameters: the rotational Reynolds number Reϕ; Rossby number Ro; the buoyancy parameter βΔT; and compressibility parameter χ. These are defined, respectively, as follows:
(2)
(3)
(4)
(5)

where Ω is the disk rotational speed, b is the cavity outer radius, W is the average throughflow velocity, Tsh is the temperature of the shroud, and Tf is the inlet throughflow temperature. Other symbols are defined in the nomenclature. Note that the Grashof number (Gr=Reϕ2βΔT) is often used as a governing parameter. In aero-engines, Gr is O(1013) creating a challenge for computational fluid dynamics to achieve accurate solutions.

Closed Cavity (η=0):

The closed cavity is a canonical case that reduces the complexity of the system, isolating the buoyancy-induced phenomena from the interaction with the axial throughflow. The Aachen group measured shroud heat transfer correlations in closed cavities [2], with adiabatic disks, providing evidence of laminar free convection. One such closed cavity (a/b=0.521 and s/b=0.5) has been the subject of several computational studies. Pitz et al. [3] conducted large-eddy simulation (LES) of flow in this geometry, finding counter-rotating vortices consistent with Rayleigh-Bénard convection and disk boundary layer profiles similar to that of laminar Ekman layers, consistent with the assumptions used by Owen and Tang [4]. Saini and Sandberg [5,6] used LES computations to show the number of structure pairs varied with time, and that compressibility effects at high rotational Reynolds numbers suppressed the formation of these structures and the shroud heat transfer, in agreement with the theoretical findings by Tang and Owen [7].

Flow structures, as well as disk and shroud heat transfer, were measured in a closed cavity with a/b=0.454 and s/b=0.167 using the Bath Compressor Cavity Rig [8]. Unsteady pressure measurements showed 3-4 vortex pairs and structure slip-to-rotation ratios of less than 1%. A plume model was developed for the prediction of disk and core temperatures by Tang and Owen [9], assuming convective heat transfer via hot and cold plumes between the vortex pairs and conductive disk heat transfer via laminar Ekman layers. The predictions were validated by Lock et al. [10] where a correlation for shroud heat transfer was established using heat flux gauge measurements and experimentally derived local core temperatures. The plume model was further developed for transient operating conditions by Nicholas et al. [11]. Unsteady Reynolds-Average Simulations in a conjugate heat transfer model from Parry et al. [12] were in good agreement with both the experiments and the model.

Open Cavity (η=1):

For cases with η=1, the disk cobs are removed and the cavity is fully open to interaction with the axial throughflow. Farthing et al. [13] conducted an experimental study of flow structure in such cavities with a/b0.1 and 0.124<s/b<0.532 using laser Doppler anemometry. They showed that the average velocity of the buoyancy-driven structures could be up to 10% slower than that of the disks and that there was a peak in this structure slip for a given Ro. They also demonstrated the existence of a toroidal vortex that significantly affected the flow structure. Bohn et al. [14] measured flow structure in a fully open cavity with a/b=0.3 and s/b=0.2, showing one pair of vortices rotating at 88–90% of the disk speed. Disk heat transfer was measured by both Bohn et al. [14] and Farthing et al. [15]; however, no consistent trends on the effects of Ro and s/b were observed. LES results from Pitz et al. [16] with a fully open cavity of a/b=0.521 and s/b=0.5 and showed through flow penetration into the cavity increased with Rez. The disk boundary layer thickness was again consistent with that of laminar Ekman layers.

Open Cavity (0<η<1):

In practice, aero-engine cavities feature 0<η<1 due to thick cobs at the inner radii of the disks (however, some industrial gas turbines incorporate virtually closed cavities (η0) with small clearances for leakage flow). Long et al. [17] and Long and Childs [18] measured velocities and shroud heat transfer in an open cavity with a/b=0.318, s/b=0.195, and η=0.601. The velocity measurements showed two flow regions in the cavity: at low radii the flow was strongly affected by the radial jet induced by axial throughflow; at high radii the flow was governed by buoyancy effects. The frequency spectrum of the tangential velocity showed two pairs of structures. Atkins and Kanjirakkad [19] presented the effects of operating parameters on disk temperature measurements in the same cavity, from which the calculated disk heat transfer by Tang et al. [20] showed suppressed buoyancy effects at high Reϕ due to compressibility.

Heat transfer and flow structure in an open cavity with a/b=0.454, s/b=0.167, and η=0.650 were investigated in the Bath Compressor Cavity Rig [2123]. A single pair of counter-rotating structures was observed slipping at 15% in the open cavity, much higher than the equivalent slip of 1% per structure in the closed cavity. This open cavity slip was a function of Ro in the open cavity, and the strength of buoyancy, βΔT, in both cavity geometries. At low Ro there was experimental evidence of egress of cavity flow to the axial throughflow in both the upstream and downstream direction. Such reversal flow recirculates heated air from the cavity and reduces the overall heat transfer in the cavity. Pernak et al. [24] have defined subcritical and supercritical flow regimes: in the former, reversal flow is prominent; in the latter the toroidal vortex in the cob region is dominant. A wall-modeled large-eddy simulation (WMLES) study from Gao and Chew [25] using Bath data and geometry for Ro >0.4 showed that axial throughflow was entrained radially into the cavity via the cold plumes, and hot air was expelled into the throughflow via the hot plumes. They also calculated flow reversal, which decreased as Ro increased with a corresponding reduction in cavity-throughflow exchange. The plume model for disk temperature prediction was applied to this open cavity by Nicholas et al. [26], determining experimentally-derived correlations for reversal flow and the mass exchange.

A recent study by Fischer and Puttock-Brown [27] gave further evidence for the toroidal vortex in the Sussex rig for 0.11<Ro<3.24, with time-averaged radial velocity measurements that were the first of their kind. The authors showed the vortex developed primarily for Ro>0.2 and that its size and strength increased with Ro.

Various experimental, theoretical and computational studies conducted by different groups have implemented both open and closed cavities with varying cavity dimensions. However, there is a gap in the literature investigating the effect of varying the cob geometry on the shroud heat transfer, flow structure, core and disk temperatures. This paper is the first to study the effect of η on mass and heat exchange for compressor cavities. A combined experimental and modeling approach has been used to capture this effect using cavities with different cob separations. Section 2 introduces the Bath Compressor Cavity Rig, the cavity geometries (η=0,0.325,0.650), and the methods of data analysis. A plume model is derived in Sec. 3 to predict the effect of η on disk and core temperatures. In Secs. 4 and 5, experimental measurements of the flow structures, shroud heat transfer, and disk temperatures over a range of Ro, Gr, and Reϕ are presented, and compared with the model. Section 6 gives practical design implications and Sec. 7 summarizes the key conclusions.

2 Experimental Apparatus and Analysis

2.1 Bath Compressor Cavity Rig.

The Bath Compressor Cavity Rig is shown in Fig. 2. Details are available in Luberti et al. [8]. The test section consists of four titanium disks, with a central cavity instrumented with thermocouples across the disk radius, a shroud heat flux gauge and two unsteady pressure sensors at r/b=0.85. This data is collected in the rotating frame of reference via a telemetry system. The external surfaces of the disks in the central cavity disks are insulated with 5 mm of low-conductivity Rohacell foam, forming a quasi-adiabatic boundary condition. The axial throughflow is supplied from ambient air in the laboratory and circular heaters at the shroud (combined with a range of rotational disk speeds) provide a range of Gr and Ro. Aluminum ring attachments to the cobs allow the rig to simulate both open and closed cavity configurations. These were bolted to the inner cob surfaces to rotate with the disks.

Fig. 2
Bath compressor cavity rig test section. Cob inserts can be removed or replaced to produce the three geometries tested in this investigation: A is a fully closed cavity; B and C are open cavities. Adapted Pernak et al. [28].
Fig. 2
Bath compressor cavity rig test section. Cob inserts can be removed or replaced to produce the three geometries tested in this investigation: A is a fully closed cavity; B and C are open cavities. Adapted Pernak et al. [28].
Close modal

Figure 2 shows the three geometries, denoted as Cavity A, B, and C and characterized by η=0,0.325 and 0.650, respectively. The inner and outer radii of the disks, the axial width in the outer region, and the bore flow dimensions are consistent in all configurations. Detailed geometrical information and operating parameters are listed in Table 1. Thirteen cases are presented in this paper but a total of 189 steady-state cases were collected using the three geometries. The detailed operational parameters of the cases with temperature distributions presented in this paper are given in Table 2.

Table 1
Shroud radiusb (mm)240
Shroud diaphragm radiusb (mm)235
Hub diaphragm radiusa (mm)125
Cob inner radiusa (mm)70
Shaft radiusrs (mm)52
Rotational speedN (RPM)800-8000
Rotational Reynolds numberReϕ0.323.1×106
Mass flow ratem˙f (kg/s)0-0.15
Axial Reynolds numberRez04×104
Rossby numberRo0-2.0
Buoyancy parameterβΔT0-0.3
Grashof numberGrf02×1012
Shroud radiusb (mm)240
Shroud diaphragm radiusb (mm)235
Hub diaphragm radiusa (mm)125
Cob inner radiusa (mm)70
Shaft radiusrs (mm)52
Rotational speedN (RPM)800-8000
Rotational Reynolds numberReϕ0.323.1×106
Mass flow ratem˙f (kg/s)0-0.15
Axial Reynolds numberRez04×104
Rossby numberRo0-2.0
Buoyancy parameterβΔT0-0.3
Grashof numberGrf02×1012
Table 2

Parameter summary for cases with temperature distributions presented in this paper

Case IDReϕβΔTRoGr/1011χτsh
A-12.3×1060.250.34130.6416
A-22.3×1060.110.336.01.68.3
B-10.80×1060.260.211.70.06815
B-20.80×1060.250.411.70.06822
B-30.80×1060.260.811.70.06918
B-42.2×1060.250.41120.6441
B-52.3×1060.160.408.81.129
B-62.9×1060.230.23201.236
C-10.81×1060.270.201.70.06648
C-20.80×1060.260.401.70.07064
C-30.81×1060.250.801.60.07360
C-42.3×1060.240.39120.67122
C-52.3×1060.100.395.51.765
Case IDReϕβΔTRoGr/1011χτsh
A-12.3×1060.250.34130.6416
A-22.3×1060.110.336.01.68.3
B-10.80×1060.260.211.70.06815
B-20.80×1060.250.411.70.06822
B-30.80×1060.260.811.70.06918
B-42.2×1060.250.41120.6441
B-52.3×1060.160.408.81.129
B-62.9×1060.230.23201.236
C-10.81×1060.270.201.70.06648
C-20.80×1060.260.401.70.07064
C-30.81×1060.250.801.60.07360
C-42.3×1060.240.39120.67122
C-52.3×1060.100.395.51.765

Not shown in Fig. 2 are thermocouple rakes that measure the radial distribution of temperature in the throughflow annulus (rs<r<a—see Fig. 1), as shown in Jackson et al. [29]. These are at the inlet to the annulus, upstream and downstream of the cavity. Such measurements identify whether flow reversal (associated with cavity ingress and egress) is significant. From Ref. [29], the flow may then be separated into subcritical (Ro < 0.4) and supercritical flow regimes (Ro > 0.4): in the former, reversal flow is prominent; in the latter the toroidal vortex in the cob region is dominant. Typical sub- and supercritical throughflow temperature profiles are shown in  Appendix, Fig. 16, for both open cavity geometries B and C.

2.2 Data Analysis.

Structure Slip and Vortex Pairs.

Experimental measurements of the unsteady pressure in the rotating core of the cavity are obtained using two sensors at r/b=0.85 on the surface of the disk. As described by Jackson et al. [21,22], the time lag between pressure signals, Δtα, is calculated using the cross-correlation of the data from each pressure sensor. The slip speed of the structures is given by the following:
(6)
where α is the circumferential separation of the sensors (α=35deg). Here, Ωs is the slip speed in the relative frame of reference and thus is small relative to Ωd. Although the slip speed is measured at a single radius, it is expected to be broadly invariant with radius, evidenced by numerical simulation from Parry et al. [12]. From Eq. (6), the number of vortex pairs is given by
(7)
where fs,1 is the passing frequency of all vortex pairs, corresponding to the first peak frequency shown on a fast Fourier transform (FFT). For the FFTs featured in Sec. 4, the pressure coefficient is defined by the difference between normalized instantaneous and mean sample pressures
(8)

Plume Mass Flow Rate.

The principal assumption in the Tang and Owen [9] plume model is that the plumes of radially flowing fluid act as the primary mechanism for heat transfer from and to the hot shroud and cool axial throughflow (or cool hub, in the case of a closed cavity). The pairs of anticyclonic and cyclonic vortices provide the necessary Coriolis forces for radial flow in the core. The pressure difference between the centers of the vortices, Δp, is determined by fitting sinusoidal curves to the experimentally measured unsteady pressure. The plume mass flowrate was linked to the resultant nondimensional circumferential pressure variation, CΔp, as follows:
(9)
where ψp is the nondimensional plume mass flowrate and the pressure coefficient CΔp is given by
(10)

where Δp is the pressure difference between the centers of the anticyclonic and cyclonic vortices.

Shroud Heat Transfer.

The heat flux gauge at the surface of the inner shroud was calibrated by Pountney et al. [30], with the measurements corrected for radiation [31]. This shroud heat flux is presented in nondimensional form
(11)

where kd and td are the thermal conductivity and thickness of the disk.

3 Plume Model for Buoyancy-Induced Heat Transfer in Different Cavities

Theoretical models have been developed separately for closed and open cavities [11,26]. This section presents an integrated model to capture the effect of cob separation ratio η on the mass exchange with the axial throughflow and consequences for buoyancy-induced flow and heat transfer.

3.1 Flow Structure.

Figure 3 illustrates the flow structure in the cavities: laminar Ekman layers with radial outflow, shroud and hub free convection layers, radial movement of hot and cold plumes, and a toroidal vortex at low radius. These illustrations are schematic as the flow is unsteady, unstable and three-dimensional. The cavity can be divided into an inner region dominated by flow exchange between the cavity and axial flows, and an outer region driven by buoyancy-induced cyclonic and anticyclonic vortices. Plumes of hot and cold fluid flow radially between these rotating vortex pairs. These are shown more clearly in Fig. 4, where the varying cob geometries have a significant effect on this flow structure. Shown in more detail in Sec. 4, increasing η reduces the number of vortex pairs and increases the significance of the exchange and reversal mass flows, ψex and ψr, respectively.

Fig. 3
Three configurations with the circumferentially and time-averaged flow structure, which is discussed further in Fig. 4
Fig. 3
Three configurations with the circumferentially and time-averaged flow structure, which is discussed further in Fig. 4
Close modal
Fig. 4
Flow structure of the three configurations: (a) η = 0, (b)η = 0.325, and (c) η = 0.650
Fig. 4
Flow structure of the three configurations: (a) η = 0, (b)η = 0.325, and (c) η = 0.650
Close modal
As discussed in Sec. 2.2, the mass flowrate of the plumes can be determined using experimental measurements of unsteady pressure. Tang and Owen [9] correlated the nondimensional flowrate with Grashof number
(12)
where Grsh is defined using the experimentally derived local core air temperature such that
(13)

For all cavities, the flows on the disks and shroud surfaces are governed by laminar Ekman layers and a free convection layer, respectively. The source flow of the cold plumes characterizes the interface between the inner and outer regions. For closed cavities (η=0), the recirculated flow in the outer region is cooled by the cobs (the hub) and is the source of the fluid in the cold plumes. For open cavities (η>0), a fraction of the cold plume flow is provided by entrainment (or ingress-egress exchange) of cold fluid from the axial throughflow. The plume mass flowrate is an upper limit for the entrainment/exchange flowrate.

The entrainment for open cavities is governed by the Rossby number. At high Ro (>0.4), the impingement of axial throughflow on the downstream disk surface forms a toroidal vortex in the inner region, inhibiting entrainment of cold fluid to the cavity. The exchange mass flowrate is quantified in nondimensional form by
(14)

At high Ro, where the axial throughflow impingement is strong even for cases with η=1, a part of the cold plume flow in the outer region is supplied by recirculation. This is illustrated in Fig. 3. At low Ro (<0.4), the toroidal vortex is suppressed and a greater proportion of the plume flow is provided by entrainment.

The effect of Ro on the exchange flowrate can be modeled as follows:
(15)

For the closed cavity (η=0), ψex=0. For the open cavities, the empirical exchange-flow coefficients D1, D2, and D3 are a function of η. The values for the two geometries B and C (η=0.325 and 0.650) were determined from experiments and presented in Sec. 5. The constant D1 will increase from cavity B to cavity C, as η increases.

The interaction between cavity and axial throughflow is affected by flow reversal at low Ro. At Ro=0, the reversal flowrate approaches the exchange mass flowrate (ψrψex), and at high Ro, flow reversal disappears due to strong axial flows (ψr0). Experimentally-derived correlations for the reversal flowrate with Ro and βΔT are presented in Sec. 5 in the following form:
(16)

3.2 Heat Transfer.

Analogous with free convection on a hot plate in a gravitational field, the heat transfer on the shroud and hub surfaces may be correlated to a Grashof number with the gravitational acceleration replaced by centripetal acceleration (typically 104g in a compressor). A general form of the correlation is
(17)
(18)
where the shroud and hub Nusselt numbers, Nush and Nuhb, and the hub Grashof number Grhb are defined as
(19)
(20)
(21)

Pr is the Prandtl number and other symbols are defined in the nomenclature.

The disk heat transfer is driven by conduction in laminar Ekman layers [10,26], with a corresponding Nusselt number calculated using
(22)
where δ* is the effective Ekman layer thickness and Nud is defined as
(23)

3.3 Disk and Core Temperatures.

Nondimensional temperatures are defined by
(24)
The core temperature is assumed to be the average of the hot and cold plume temperatures. The radial distribution of cold plume temperature in the outer region (see Figs. 3 and 4) is derived using the energy equation [9]
(25)
where χ is the compressibility parameter
(26)
In Eq. (25), θp̂,a is the cold plume temperature at the lowest radius of the outer region, the integral term calculates heat transfer to and from the Ekman layers and the disk, and the term with χ gives the contribution of compressibility. Note that this outer cold plume temperature distribution is based on the plume mass flowrate, ψp. In the inner region the radial distribution of cold plume temperature is driven by the exchange mass flowrate, ψex
(27)

where Φ is the angle (as a function of radius) between the upwards radial direction and the elemental disk surface area, which is zero except for the hub fillet region.

The above requires initial values at the inlet to the inner and outer regions, r=a and r=a, respectively. When there is no reversal flow, θp̂,a=0, i.e., the cold plume and inlet throughflow temperatures are equal. When reversal flow is significant, the temperature of this exchange flow is increased due to enthalpy exchange with the hot plume temperature, θpˇ. This results in the following:
(28)
At the inlet to the outer region, the cold plume temperature is increased by mixing with the recirculated flow in the outer region, which results in a step change in the cold plume temperature. It follows that
(29)
where τhb is the nondimensional hub heat flux, Ahb is the hub area, and yd is the nondimensional disk thickness defined by
(30)
τhb is calculated using
(31)

Thus, when the cavity is fully-open, the hub area Ahb is reduced to 0 and the central right-hand side term in Eq. (29) reduces to zero. This term calculates contribution of convective heat transfer from the hub to the cold plume temperature in the outer region. The right most term describes the contribution of disk heat transfer and compressibilty to the outer cold plume temperature, as a function of both the plume and exchange mass flows. If the cavity is fully closed (cavity A, η=0) then the inner region is replaced by cobs and the hub free convection layer, so this equation reduces to just the hub heat flux term.

The radial distribution of hot plume temperature can be calculated again using the energy equation. In the outer region,
(32)
In the inner region
(33)
The hot plume temperature at r=b can be computed assuming all the heat transferred from the shroud increases the plume temperatures. Hence, it follows
(34)

where Ash is the shroud area and τsh is the nondimensional shroud heat flux. In Eq. (34), θp̂,b is the cold plume temperature at r=b, and the right most term is the temperature difference between the plumes caused by convective heat transfer from the shroud.

The core temperature is then computed by
(35)
Calculation of the disk temperature is a conjugate problem where the convective heat transfer in the cavity and the conduction within the disk need to be coupled. The conduction within the disk can be solved by
(36)
where nondimensional axial distance is defined by
(37)
with the boundary conditions on the disk surface determined by the core temperature via nondimensional disk heat flux,
(38)
The boundary conditions used to solve Eq. (36) are given as
(39)

where θl is the nondimensional temperature of the core in the cavities upstream and downstream of the test section. As discussed in Sec. 2, these adjacent cavities were isolated from the axial throughflow (Fig. 2) and thermal-insulating foam created a quasi-adiabatic boundary condition on the back surfaces of the disks. Temperature boundary conditions were applied at the outer radius of the disk diaphragm and the inner radius of the cobs. Heat flux boundary conditions were applied on the disk surfaces using the characteristics of laminar Ekman layers, [11,26]. Note that due to the conjugate nature of the problem, the temperatures of the core and disk were solved iteratively. The iterative process was repeated until the change of the disk temperature Δθd<105. Depending on the initial guess, this typically only took 5-6 iterations until convergence. The calculation of the disk and core temperatures for one experimental case takes approximately two seconds using a standard laptop.

The calculation of θd was relatively insensitive to the air temperatures in the upstream and downstream cavities, θl, which were isolated from the axial throughflow - see Fig. 2. A 10% change in θl produced <1% change in θd. Note that the disk temperatures at r=b and r=a were fixed with measured values. Ideally, a correlation for the cooling on the inner cob surface can be used to predict the temperature at r=a, and the temperature at r=b can be determined by the heat transfer on the outer shroud surface and adjacent disks.

4 Experimental Results—Flow Structure and Shroud Heat Transfer

4.1 Flow Structure, Slip, and Vortex Pairs.

As discussed in Sec. 2.2, cross-correlation of the two pressure signals provides the slipping speed (Ωs) of the counter-rotating structures. Figure 5 shows the variation of slipping speed with different Ro for all three cavities at βΔT=0.25. Note that for the closed cavity (Cavity A), the interaction between the cavity and throughflow is prevented, hence there is no direct effect of Ro on the cavity flow structure.

Fig. 5
Slip of rotating structures (accounting for the number of pairs) against Rossby number for βΔT=0.25
Fig. 5
Slip of rotating structures (accounting for the number of pairs) against Rossby number for βΔT=0.25
Close modal

Consistent with the findings in Jackson et al. [21] and Nicholas et al. [26], there is a critical Ro for which there is a peak in structure slip, showing maximum momentum exchange between the open cavity and throughflow. This peak is a combined result of the reversal flow in the axial throughflow and the toroidal vortex generated by the throughflow in the inner region. As η is halved from Cavity C to B, the magnitude of Ωs decreased from 1015% to 24%, indicating an nonlinear reduction in the exchange mass flowrate. In the closed cavity with no momentum exchange, Ωs<1% for all cases and no distinct peak with Ro is observed (as there is no exchange mass flowrate for closed cavities, the parameter Ro for Cavity A is merely an indication of the cooling on the inner cob surface).

Figure 6 shows the fast Fourier transform (FFT) for all three geometries for Ro=0.2, βΔT=0.25 and Reϕ=8×105. The magnitude of Cp denotes the approximate strength of the rotating structures, and this is maximum for Cavity C; here the open area at the cobs is largest and momentum exchange with the axial throughflow is greatest. The frequency of the peak in Cp is the passing frequency of all vortex pairs, fs,1. Equation (7) reveals a single pair for Cavity C, with the second peak, at fs,2=2×fs,1, corresponding to the asymmetry of the two vortices, as detailed in Ref. [23]. There are three, virtually-symmetric vortex pairs for Cavity A. While these rotating vortices are inherently unsteady, cavities A and C exhibit a regular set of pairs. The intermediate cavity (B) displays an aperiodic behavior. The FFT shows two equal peaks between 0.062<f/fd<0.074, corresponding to either two or three vortex pairs, as illustrated by the spectrogram in Fig. 7.

Fig. 6
Example FFT comparison for the three geometries, at Reϕ=8×105, Ro=0.2, and βΔT=0.25
Fig. 6
Example FFT comparison for the three geometries, at Reϕ=8×105, Ro=0.2, and βΔT=0.25
Close modal
Fig. 7
Spectogramd for steady-state temperatures revealing unsteady flow structure for cavity configurations (a), (b), and (c), respectively. Reϕ=0.80×106, βΔT=0.25, and Ro=0.2. (a) η = 0, (b) η = 0.325, and (c) η = 0.650.
Fig. 7
Spectogramd for steady-state temperatures revealing unsteady flow structure for cavity configurations (a), (b), and (c), respectively. Reϕ=0.80×106, βΔT=0.25, and Ro=0.2. (a) η = 0, (b) η = 0.325, and (c) η = 0.650.
Close modal

As shown in Fig. 7, the closed cavity features a peak frequency at f/fd=0.022, with some noise at lower frequencies and three pairs of structures. Cavity C (that with the largest opening at the cobs) shows a constant peak frequency of f/fd=0.13 with some unsteadiness and a higher level of overall noise caused by the throughflow interaction. Here there is a single pair of structures with the second peak (fs,2/fd=0.26) resulting from vortex asymmetry. Cavity B features two distinct bands of frequency peaks at f/fd 0.05 and 0.07, and no clear presence of vortex asymmetry. Cross-correlation of a variety of samples from this case shows that the number of structure pairs are changing frequently in an aperiodic nature, along with a slight change in slip per structure. The upper band corresponds to three pairs of structures and with slip Ωs/Ωd=2.3% per structure, and the lower band to two pairs of structures with slip Ωs/Ωd=2.5%.

A summary of the number of vortex pairs (given by the ratio peak frequency, fs,1, to structure slip, Ωs) for all cavities with varying Reϕ is given in Fig. 8. For Cavity C there is one pair of structures found from cross-correlation and FFTs of the dual pressure signals. For the closed cavity A, there is a transition from three to four pairs for Reϕ2.2×106. For Cavity B, the number of structure pairs is between two and three. The reason for a noninteger value is that the results are averaged across 500 disk revolutions to produce a mean result from the aperiodic structures.

Fig. 8
Ratio of peak frequency to structure slip, indicating number of pairs of rotating structures. For varying Reϕ and for the three configurations at βΔT=0.25 across the entire Ro range.
Fig. 8
Ratio of peak frequency to structure slip, indicating number of pairs of rotating structures. For varying Reϕ and for the three configurations at βΔT=0.25 across the entire Ro range.
Close modal

4.2 Plume Mass Flow Rate and Shroud Heat Transfer.

Plume mass flow rates, ψp, were derived from the measurements of unsteady pressure using Eq. (9). The shroud Grashof number Grsh (Eq. (13)) was calculated using the measured shroud temperature and the experimentally-derived core temperature Tc,b from the disk temperature measurements in the outer region and the heat transfer coefficients from the conductive Ekman layer. Note there is a distinction between the experimentally-derived core temperatures and the theoretical- predicted core temperatures from the model in Sec. 5. The former involves solving the inverse problem by matching resultant disk temperatures with experimental values. The predicted values come from the correlations for heat transfer and mass flow.

Figure 9 shows the correlations for ψp in the three cavities. As η increases, the plume mass flowrate decreases at constant Grsh, which is attributed to the increase in the slip speed. This increased slip effectively reduces buoyancy forces and suppresses radial movement. The magnitudes for Cavity A and B are similar as the slip in both cases are similar. Equation (12) leads to the following correlations for the three cavities
(40)
(41)
(42)
Fig. 9
Correlation between the nondimensional plume mass flowrate and shroud Grashof number for all three geometries. Solid lines denote correlations from fitted maximum likelihood estimation (MLE) of all viable points.
Fig. 9
Correlation between the nondimensional plume mass flowrate and shroud Grashof number for all three geometries. Solid lines denote correlations from fitted maximum likelihood estimation (MLE) of all viable points.
Close modal
The shroud Nusselt number Nush was determined using the measured shroud heat flux corrected for radiation [31]. Figure 10 shows Nush for all cavities. In contrast to the plume mass flowrate, for a given Grsh, the shroud heat transfer is enhanced as η increases. This can be explained by the additional forced convection component in open cavities as a result of the increased momentum exchange and resultant increase in slip of the cavity flow. Equation (17) gives the following correlations for the three cavities
(43)
(44)
(45)
Fig. 10
Correlation between the shroud Nusselt and Grashof numbers for all three geometries. Solid lines denote correlations from fitted MLE of all viable points.
Fig. 10
Correlation between the shroud Nusselt and Grashof numbers for all three geometries. Solid lines denote correlations from fitted MLE of all viable points.
Close modal

Both the plume mass flowrate and shroud heat transfer correlations are used in the plume model to determine the magnitude and distribution of core and disk temperatures, as discussed in the next section.

5 Prediction of Mass and Heat Transfer and Cavity Temperatures

In the closed cavity the disk and core temperatures were calculated with zero exchange mass flowrate. For both open cavities, the exchange mass flow rates were required to determine the heat exchange and, in turn, the disk and core temperatures. As discussed in Sec. 3.1, the exchange flow is subject to the reversal of axial throughflow at low Ro and the toroidal vortex at high Ro. The reversal flowrate was calculated using Eq. (16). Consistent with results in Ref. [26], E=0.24 was used for both open cavities. Correlations for the exchange-plume mass flow ratio were obtained for both open cavities by regression of the experimentally-derived values, and illustrated in Fig. 11 
(46)
(47)
Fig. 11
Correlations of experimentally-derived exchange mass flowrate with Ro and βΔT for cavities B and C
Fig. 11
Correlations of experimentally-derived exchange mass flowrate with Ro and βΔT for cavities B and C
Close modal

The predicted cavity temperatures at Reϕ=2.3×106, βΔT=0.25 and Ro=0.4 are shown in Fig. 12 (though the Rossby number is not the driving parameter for closed cavity heat transfer, it is presented to show consistent bore cooling conditions with the open cavity cases). Symbols denote disk temperature measurements (averaged across both upstream and downstream disks). Solid lines denote disk temperature predictions from the plume model and dashed lines core temperature predictions. These predictions follow the procedure described in Sec. 3.3, using correlations for shroud and hub heat transfer, plume, reversal, and exchange mass flow, and the temperature and heat flux boundary conditions. In Cavity A, the temperature predictions are limited to the diaphragm section of the test geometry (0.52<x<0.98, the outer region). For the open cavities, this prediction extends to the inner or cob region; here there is a step change in core temperature as the radial flow is driven by plume mass flow in the outer region and driven by exchange mass flow in the inner region, where ψex<ψp. There is an increase core temperature with radius caused by the compressibility effect at large Reϕ. Note that there is consistent agreement between the measured and modeled disk temperatures for all three geometries.

Fig. 12
Effect of cob geometry on the radial distribution of disk and core temperatures for fixed operating conditions, Reϕ=2.3×106, βΔT=0.25, Ro =0.4. Solid symbols denote experimental measurements, solid lines the predicted disk temperatures, and dashed lines the predicted core temperatures.
Fig. 12
Effect of cob geometry on the radial distribution of disk and core temperatures for fixed operating conditions, Reϕ=2.3×106, βΔT=0.25, Ro =0.4. Solid symbols denote experimental measurements, solid lines the predicted disk temperatures, and dashed lines the predicted core temperatures.
Close modal

For η=0 (Cavity A), the core temperature is lower than the disk temperature (θc<θd) at high radii and higher at low radii, and the crossover radius (where θc=θd) is near x=0.82. This is consistent with results from Lock et al. [10]. The crossover radius is closer to the shroud than the hub due to the larger shroud surface area and heat transfer.

As η increases, the mass and heat transfer between the cavity and throughflow is increased, hence θd, θc and the cross-radius decreases. Despite Cavity B being open to interaction with the throughflow, the distributions of disk temperature were similar to those of Cavity A. This was caused by the reduced exchange mass flowrate relative to Cavity C, equivalent to just 510% of the plume mass flow. There is a step increase of the core temperature from the inner to outer region for cavities B and C. This is caused by the mixing of the cold plume flow with the hot recirculated flow in the outer region.

Figure 13 is similar to Fig. 12, other than reduced βΔT. Consider cavity C, with the largest η. Equations (41) and (44) show a reduction in βΔT leads to reduction in Grashof number and ψp, resulting in a decrease in ψex and, in turn, an increase in the core temperature θc. In turn, there is a reduction in the disk temperature gradient in the outer region. However, for cavities A and B, the exchange flowrate is very low or zero. Thus, there is a relatively minor effect from βΔT on the overall distributions of nondimensional temperature, except a reduction in cob temperature. This is due to weakened convection from the shroud as a result of the reduced Grashof number (see Eq. (17)). Relative to Fig. 12 there is an increase in the gradient of core temperature due to the reduced βΔT, which causes an increase in the compressibility parameter χ (see Eq. (26)). This is a result of the dimensional core temperature gradient being proportionately larger due to the lower ΔT=TshTf value.

Fig. 13
Radial distribution of disk and core temperatures for all geometries at low Gr. Reϕ=2.3×106 and Ro =0.4.
Fig. 13
Radial distribution of disk and core temperatures for all geometries at low Gr. Reϕ=2.3×106 and Ro =0.4.
Close modal

Figure 14 shows the impact of Reϕ on temperatures for cavity B. There is a reduction in core temperature with increased Reϕ, which is due to the increase in Grashof number, plume mass flow, and mass and heat transfer between the cavity and throughflow. Moreover, increasing Reϕ increases the compressibility parameter, χ, and so increases the radial gradient of core temperature, as observed in both other cavity configurations [10,26].

Fig. 14
Radial distribution of temperatures for cavity B at different disk rotational speeds, βΔT=0.25, Ro =0.2
Fig. 14
Radial distribution of temperatures for cavity B at different disk rotational speeds, βΔT=0.25, Ro =0.2
Close modal

Figure 15 illustrates the impact of changing from subcritical to supercritical Rossby number for cavities B and C. The former is the reversal flow regime where Ro<0.4 and the latter the toroidal vortex regime where Ro>0.4. As Ro increases to 0.4, there is a decrease in both θd and θc due to the reduction in reversal flow, ψr. From Figure 11, ψex is approximately constant for Ro<0.4, but ψr decreases to zero, see Eq. (14). For Ro>0.4, the toroidal vortex reduces ψex, causing a slight increase in θd and θc. This trend in θc is consistent with that of structure slip in Fig. 5, showing a peak in mass and heat transfer between the cavity and throughflow, as well as the peak in shroud heat transfer for Ro=0.4 shown in Table 2.

Fig. 15
Radial distribution of temperatures for (a) cavity B and (b) cavity C for different Rossby numbers at βΔT=0.25, Reϕ=0.80×106
Fig. 15
Radial distribution of temperatures for (a) cavity B and (b) cavity C for different Rossby numbers at βΔT=0.25, Reϕ=0.80×106
Close modal

6 Practical Design Implications

This research has illustrated the nonlinear effect of varying the cob separation on the heat transfer and flow structure in aero-engine compressor cavities. The radial variation of disk temperature is influenced significantly by this cob separation, as will be the thermal stresses and expansion of the compressor rotor. These, in turn, will impact tip clearances, engine operating life and efficiency. The model presented here provides expedient, reduced-order solutions to the complex conjugate heat transfer problem. The methodology is specifically intended for incorporation into practical thermo-mechanical engine design codes.

Engine designers have the challenge of balancing component life with a system-level optimization. From one perspective, enhanced heat transfer effectively cools the disks reduces tip clearances; however, this causes a maximum enthalpy exchange with the axial throughflow, reducing its cooling effectiveness in the later stages of the engine and increases thermal stresses. A compromise governed by the cavity-bore flow interaction is required. The model (validated here by experimental data) provides a quantitative prediction of both the disk temperatures and throughflow enthalpy exchange, better informing iterative engine design decisions.

7 Conclusions

This paper provides the first study of the effect of cob separation on mass and heat exchange in rotating compressor cavities. A new model has been presented that is able to predict the interaction between the cavity and the axial throughflow, and consequently the radial distribution of disk and fluid-core temperatures. The model was validated using detailed experimental measurements from the Bath Compressor Cavity Rig, which has the unique capability of quantifying disk temperature and shroud heat flux alongside unsteady pressure signals in the rotating cavity. A range of disk-cob spacings was investigated over a range of engine representative conditions.

Measurements of pressure from fast-response sensors on the rotating disks captured unsteady cyclonic and anticyclonic vortex pairs typical of buoyancy-induced flow. The number of structures and their slip relative to the disk depended on the cob separation, with different degrees of ingress/egress to/from the cavity. Specific cases featured fundamentally aperiodic (chaotic) features, even at steady-state conditions of Rossby, Grashof and Reynolds numbers.

Correlations for shroud Nusselt number and radial mass flowrate in the cavity were generated across a wide range of Grashof numbers. The convective heat transfer at the shroud was highest for the largest cob separation, with increased cavity-throughflow momentum exchange and ingress of cold fluid. The rate of radial mass flow reduced as the cob separation increased, with larger slip (relative to the disk) of the rotating structures at increased cavity-throughflow interaction. Correlations were generated for mass exchange between the cavity and axial throughflow, showing a nonlinear reduction in fluid entrainment with reduced cob separation.

The model was able to predict the mass and heat transfer to/from the cavities, together with the radial distributions of core and disk temperatures. Excellent agreement with experiments was achieved for 189 cases that spanned a range of Rossby (Ro), Grashof and Reynolds numbers. There is a critical Ro near 0.4 where heat transfer and core slip are maximized, and core and disk temperatures are minimized. This was shown to be a result of two phenomena: flow reversal at low Ro and a toroidal vortex at high Ro.

The physics-based model developed here is able to provide a reduced-order method to support the prediction of thermal stresses and blade-tip clearance in high-pressure compressors. Appropriate for thermo-mechanical design in industry, the model was created to inform the design of next generation high pressure-ratio aeroengines that require ever greater improvements to efficiency and accurate tip-clearance prediction.

Acknowledgment

The research presented in this paper was supported by the UK Engineering and Physical Sciences Research Council and in collaboration with Rolls-Royce plc and the University of Surrey, under the Grant No. EP/P003702/1. The authors are especially grateful for the support of Jake Williams and the approval from Rolls-Royce to publish the work.

Funding Data

  • Engineering and Physical Sciences Research Council (Award No. EP/P003702/1; Funder ID: 10.13039/501100000266).

Data Availability Statement

The authors attest that all data for this study are included in the paper.

Nomenclature

A =

surface area (m2)

a =

inner radius of cavity (cob) (m)

a =

inner diaphragm radius (m)

b =

outer diaphragm radius (m)

b =

outer radius of cavity (shroud) (m)

C =

plume mass flow correlation coefficient

cp =

specific heat capacity (J/(kgK))

dh =

hydraulic diameter (=2(ars)) (m)

D1,D2,D3 =

exchange flow coefficients

E =

reversal flow correlation coefficient

F =

shroud/hub heat transfer correlation coefficient

fs =

rotational frequency of the flow structure (Hz)

G =

cavity aspect ratio (=s/b)

h =

heat transfer coefficient (W/(m2K))

k =

thermal conductivity of air (W/(mK))

kd =

thermal conductivity of disk (W/(mK))

l =

empirical exchange flow exponent

m˙ =

mass flow rate (kg/s)

m =

plume mass flow coefficient exponent

n =

number of vortex pairs

N =

rotational speed of disks (RPM)

p =

static pressure (Pa)

p¯ =

mean static pressure (Pa)

q =

heat flux (W/m2)

r =

radius (m)

R =

specific gas constant (J/(kgK))

s =

cavity width at diaphragm (m)

s =

cob separation (m)

t =

thickness (m)

tα =

time lag between pressure sensors (s)

T =

temperature (K)

W =

throughflow velocity (m/s)

x =

non-dimensional radial location

y =

non-dimensional axial distance

z =

axial distance (m)

α =

circumferential angle between two pressure sensors (rad)

β =

volume expansion coefficient (K1)

γ =

ratio of specific heats

δ =

Ekman layer thickness (m)

η =

cob separation ratio (=s/s)

θ =

non-dimensional temperature (=(TTf)/(TshTf))

μ =

dynamic viscosity (kg/(ms))

ρ =

density (kg/m3)

τ =

non-dimensional heat flux (=b2q/(kdtd(TshTf)))

χ =

compressibility parameter (=Ma2/βΔT)

ψ =

non-dimensional mass flow rate (=m˙/(μs))

Φ =

angle between disk surface and radial direction (rad)

Ω =

angular velocity (rad/s)

Cp =

coefficient of pressure for FFT (=(pp¯)/(2ρfΩd2b2))

CΔp =

coefficient of pressure difference (=Δp/(ρfΩd2b2))

Gr =

Grashof number (=Reϕ2βΔT)

Ma =

rotational Mach number (=Ωb/γR(Ta+Tb))

Nu =

Nusselt number (=hb/k)

Pr =

Prandtl number (=Cpμ/k)

Ra =

Rayleigh number (=PrGr)

Ro =

Rossby number (=W/(Ωa)

Reϕ =

Rotational Reynolds number (ρΩb2)/(μ))

Rez =

Axial Reynolds number (ρWdh/(μ))

βΔT =

Buoyancy parameter (=(Tsh+Tf)/Tf)

a =

value at the inner radius of the cavity

a =

value at the inner radius of the diaphragm

b =

value at the outer radius of the cavity

b =

value at the outer radius of the diaphragm

c =

value in the fluid core

d =

value on the disk surface

E =

value in Ekman layers

ex =

value in exchange mass flow

f =

value in the axial cooling flow

hb =

value on hub surface

l =

value disk back (Rohacell) surface

p =

average value in radial plumes

pˇ,p̂ =

values in the hot and cold plumes

r =

value in reverse mass flow

s =

value of structures

sh =

value on internal shroud surface

ϕ, r, z =

circumferential, radial and axial direction

Appendix: Throughflow Temperature Plots

Figure 16 demonstrates the effect of reversal flow for open cavities, where the temperature of the throughflow just upstream of the cavity is hotter than the inlet temperature for low Ro. There is also a reduction in the magnitude of reversal flow that occurs for reduced η due to the reduced ψex. For higher Ro, the effect disappears and the upstream throughflow temperatures are approximately constant. The temperature increase downstream of the cavity is also reduced, as throughflow-cavity interaction diminishes with Ro as the toroidal vortex strength increases and ψex reduces.

Fig. 16
Temperatures for both open cavities at different axial locations in the throughflow demonstrating the effects of reversal flow at different Rossby numbers: (a) Ro =0.1 and (b) Ro =0.8
Fig. 16
Temperatures for both open cavities at different axial locations in the throughflow demonstrating the effects of reversal flow at different Rossby numbers: (a) Ro =0.1 and (b) Ro =0.8
Close modal

References

1.
Vullo
,
V.
, and
Vivio
,
F.
,
2013
,
Rotors: Stress Analysis and Design
,
Springer
, Berlin.
2.
Bohn
,
D.
,
Deuker
,
E.
,
Emunds
,
R.
, and
Gorzelitz
,
V.
,
1995
, “
Experimental and Theoretical Investigations of Heat Transfer in Closed Gas Filled Rotating Annuli
,”
ASME J. Turbomach.
,
117
(
1
), pp.
175
183
.10.1115/1.2835635
3.
Pitz
,
D. B.
,
Chew
,
J. W.
, and
Marxen
,
O.
,
2019
, “
Large-Eddy Simulation of Buoyancy-Induced Flow in a Sealed Rotating Cavity
,”
ASME J. Eng. Gas Turbines Power
,
141
(
2
), p. 021020.10.1115/1.4041113
4.
Owen
,
J. M.
, and
Tang
,
H.
,
2015
, “
Theoretical Model of Buoyancy-Induced Flow in Rotating Cavities
,”
ASME J. Turbomach.
,
137
(
11
), p.
111005
.10.1115/1.4031353
5.
Saini
,
D.
, and
Sandberg
,
R. D.
,
2020
, “
Simulations of Compressibility Effects in Centrifugal Buoyancy-Induced Flow in a Closed Rotating Cavity
,”
Int. J. Heat Fluid Flow
,
85
(
10
), p.
108656
.10.1016/j.ijheatfluidflow.2020.108656
6.
Saini
,
D.
, and
Sandberg
,
R. D.
,
2021
, “
Large-Eddy Simulations of High Rossby Number Flow in the High-Pressure Compressor Inter-Disk Cavity
,”
ASME J. Turbomach.
,
143
(
11
), p.
111002
.10.1115/1.4050951
7.
Tang
,
H.
, and
Owen
,
J. M.
,
2018
, “
Theoretical Model of Buoyancy-Induced Heat Transfer in Closed Compressor Rotors
,”
ASME J. Eng. Gas Turbines Power
,
140
(
3
), p.
032605
.10.1115/1.4037926
8.
Luberti
,
D.
,
Patinios
,
M.
,
Jackson
,
R. W.
,
Tang
,
H.
,
Pountney
,
O. J.
,
Scobie
,
J. A.
,
Sangan
,
C. M.
,
Owen
,
J. M.
, and
Lock
,
G. D.
,
2021
, “
Design and Testing of a Rig to Investigate Buoyancy-Induced Heat Transfer in Aero-Engine Compressor Rotors
,”
ASME J. Eng. Gas Turbines Power
,
143
(
4
), p.
041030
.10.1115/1.4048601
9.
Tang
,
H.
, and
Owen
,
J.
,
2023
, “
Plume Model for Buoyancy-Induced Flow and Heat Transfer in Closed Rotating Cavities
,”
ASME J. Turbomach.
,
145
(
1
), p.
011005
.10.1115/1.4055449
10.
Lock
,
G. D.
,
Jackson
,
R. W.
,
Pernak
,
M.
,
Pountney
,
O. J.
,
Sangan
,
C. M.
,
Owen
,
J. M.
,
Tang
,
H.
, and
Scobie
,
J. A.
,
2023
, “
Stratified and Buoyancy-Induced Flow in Closed Compressor Rotors
,”
ASME J. Turbomach.
,
145
(
1
), p.
011008
.10.1115/1.4055448
11.
Nicholas
,
T. E. W.
,
Pernak
,
M. J.
,
Scobie
,
J. A.
,
Lock
,
G. D.
, and
Tang
,
H.
,
2023
, “
Transient Heat Transfer and Temperatures in Closed Compressor Rotors
,”
Appl. Therm. Eng.
,
230
, p.
120759
.10.1016/j.applthermaleng.2023.120759
12.
Parry
,
J.
,
Tang
,
H.
,
Scobie
,
J.
,
Lock
,
G.
, and
Carnevale
,
M.
,
2024
, “
Conjugate Modelling of a Closed co-Rotating Compressor Cavity
,”
ASME J. Eng. Gas Turbines Power
,
146
(
5
), p.
051007
.10.1115/1.4063632
13.
Farthing
,
P. R.
,
Long
,
C. A.
,
Owen
,
J. M.
, and
Pincombe
,
J. R.
,
1992
, “
Rotating Cavity With Axial Throughflow of Cooling Air - Heat Transfer
,”
ASME J. Turbomach.
,
114
(
1
), pp.
229
236
.10.1115/1.2927990
14.
Bohn
,
D. E.
,
Deutsch
,
G.
,
Simon
,
B.
, and
Burkhardt
,
C.
,
2000
, “Flow Visualisation in a Rotating Cavity With Axial Throughflow,”
ASME
Paper No.
V003T01A084
.10.1115/2000-GT-0280
15.
Farthing
,
P. R.
,
Long
,
C. A.
,
Owen
,
J. M.
, and
Pincombe
,
J. R.
,
1992
, “
Rotating Cavity With Axial Throughflow of Cooling Air: Flow Structure
,”
ASME J. Turbomach.
,
114
(
1
), pp.
237
246
.10.1115/1.2927991
16.
Pitz
,
D. B.
,
Chew
,
J. W.
, and
Marxen
,
O.
,
2019
, “
Effect of an Axial Throughflow on Buoyancy-Induced Flow in a Rotating Cavity
,”
Int. J. Heat Fluid Flow
,
80
(
12
), p.
108468
.10.1016/j.ijheatfluidflow.2019.108468
17.
Long
,
C. A.
,
Miché
,
N. D. D.
, and
Childs
,
P. R. N.
,
2007
, “
Flow Measurements Inside a Heated Multiple Rotating Cavity With Axial Throughflow
,”
Int. J. Heat Fluid Flow
,
28
(
6
), pp.
1391
1404
.10.1016/j.ijheatfluidflow.2007.04.010
18.
Long
,
C. A.
, and
Childs
,
P. R. N.
,
2007
, “
Shroud Heat Transfer Measurements Inside a Heated Multiple Rotating Cavity With Axial Throughflow
,”
Int. J. Heat Fluid Flow
,
28
(
6
), pp.
1405
1417
.10.1016/j.ijheatfluidflow.2007.04.009
19.
Atkins
,
N. R.
, and
Kanjirakkad
,
V.
,
2014
, “
Flow in a Rotating Cavity With Axial Throughflow at Engine Representative Conditions
,”
ASME
Paper No. V05CT16A041.10.1115/V05CT16A04110.1115/GT2014-27174
20.
Tang
,
H.
,
Shardlow
,
T.
, and
Owen
,
J. M.
,
2015
, “
Use of Fin Equation to Calculate Nusselt Numbers for Rotating Disks
,”
ASME J. Turbomach.
,
137
(
12
), p.
121003
.10.1115/1.4031355
21.
Jackson
,
R. W.
,
Tang
,
H.
,
Scobie
,
J. A.
,
Pountney
,
O. J.
,
Sangan
,
C. M.
,
Owen
,
J. M.
, and
Lock
,
G. D.
,
2022
, “
Unsteady Pressure Measurements in a Heated Rotating Cavity
,” ASME
J. Eng. Gas Turbines Power
,
144
(
4
), p.
041017
.10.1115/1.4053390
22.
Jackson
,
R. W.
,
Tang
,
H.
,
Scobie
,
J. A.
,
Owen
,
J. M.
, and
Lock
,
G. D.
,
2022
, “
Measurement of Heat Transfer and Flow Structures in a Closed Rotating Cavity
,”
ASME J. Eng. Gas Turbines Power
,
144
(
5
), p.
051005
.10.1115/1.4053392
23.
Jackson
,
R. W.
,
Luberti
,
D.
,
Tang
,
H.
,
Pountney
,
O. J.
,
Scobie
,
J. A.
,
Sangan
,
C. M.
,
Owen
,
J. M.
, and
Lock
,
G. D.
,
2021
, “
Measurement and Analysis of Buoyancy-Induced Heat Transfer in Aero-Engine Compressor Rotors
,”
ASME J. Eng. Gas Turbines Power
,
143
(
6
), p.
061004
.10.1115/1.4049100
24.
Pernak
,
M.
,
Nicholas
,
T. E. W.
,
Carnevale
,
M.
,
Lock
,
G. D.
,
Tang
,
H.
, and
Scobie
,
J. A.
,
2024
, “
Flow and Heat Transfer in Rotating Compressor Cavities With Inverted Shroud-Throughflow Temperature Differences
,”
ASME
Paper No. GT2023-123211.10.1115/GT2023-123211
25.
Gao
,
F.
, and
Chew
,
J. W.
,
2022
, “
Flow and Heat Transfer Mechanisms in a Rotating Compressor Cavity Under Centrifugal Buoyancy-Driven Convection
,”
ASME J. Eng. Gas Turbines Power
,
144
(
5
), p.
051010
.10.1115/1.4052649
26.
Nicholas
,
T. E. W.
,
Scobie
,
J. A.
,
Lock
,
G. D.
, and
Tang
,
H.
,
2024
, “
A Model of Mass and Heat Transfer for Discs Temperature Prediction in Open Compressor Cavities
,”
ASME J. Turbomach.
,
146
(
4
), p.
041001
.10.1115/1.4064082
27.
Fisher
,
E.
, and
Puttock-Brown
,
M. R.
,
2024
, “
Experimental Measurements of Flow-Averaged Toroidal Vortces in Buoyancy-Dominated Rotating Cavities
,”
ASME J. Eng. Gas Turbines Power
,
146
(
4
), p.
041006
.10.1115/1.4063689
28.
Pernak
,
M. J.
,
Nicholas
,
T. E. W.
,
Williams
,
J. T.
,
Jackson
,
R. W.
,
Tang
,
H.
,
Lock
,
G. D.
, and
Scobie
,
J. A.
,
2023
, “
Experimental Investigation of Transient Flow Phenomena in Rotating Compressor Cavities
,”
ASME J. Turbomach.
,
145
(
12
), p.
121005
.10.1115/1.4063507
29.
Jackson
,
R. W.
,
Tang
,
H.
,
Scobie
,
J. A.
,
Pountney
,
O. J.
,
Sangan
,
C. M.
,
Owen
,
J. M.
, and
Lock
,
G. D.
,
2021
, “
Analysis of Shroud and Disk Heat Transfer in Aero-Engine Compressor Rotors
,”
ASME J. Eng. Gas Turbines Power
,
143
(
9
), p.
091005
.10.1115/1.4050631
30.
Pountney
,
O. J.
,
Patinios
,
M.
,
Tang
,
H.
,
Luberti
,
D.
,
Sangan
,
C. M.
,
Scobie
,
J. A.
,
Owen
,
J. M.
, and
Lock
,
G. D.
,
2021
, “
Calibration of Thermopile Heat Flux Gauges Using a Physically-Based Equation
,”
Proc. Inst. Mech. Eng., Part A: J. Power Energy
,
235
(
7
), pp.
1806
1816
.10.1177/0957650920982103
31.
Tang
,
H.
, and
Owen
,
J. M.
,
2021
, “
Effect of Radiation on Heat Transfer Inside Aeroengine Compressor Rotors
,”
ASME J. Turbomach.
,
143
(
5
), p.
051005
.10.1115/1.4050114