Sei sulla pagina 1di 22

Int. Symp.

on Heat Transfer in Gas Turbine Systems 9 14 August, 2009, Antalya, Turkey

FILM COOLING SIMULATION AND CONTROL Sumanta Acharya1 Mechanical Engineering Department Turbine Innovation and Energy Research (TIER) Center Louisiana State University, Baton Rouge, LA 70803 (* Corresponding author: acharya@me.LSU.edu)
1

ABSTRACT. Numerical simulation strategies for film cooling flows ranging from time-averaged simulations based on Reynolds-averaged equations to fully unsteady simulations are presented and reviewed. Simulation results are presented for both cylindrical and shaped holes with the key emphasis in this paper on the accuracy of the averaged results and on the film-cooling flow physics. It is shown that the Reynolds averaged equations with turbulence models need anisotropy and realizability corrections to provide correct predictions for the cooling effectiveness. Spatially-filtered unsteady simulations (called large-eddy simulations) appear to provide the needed quantitative accuracy and the rich flow physics representing such flows. INTRODUCTION Turbine surfaces are exposed to high-temperature gases that exceed material limits and have to be actively cooled. Film cooling is commonly employed in the high pressure turbine. With this strategy, a row of coolant jets is injected at an angle from the airfoil surface or endwall into the heated crossflow. The role of the coolant jets is to provide effective coolant-film-coverage on the airfoil surface, and to protect the airfoil from the hot crossflow. Therefore, it is desirable to minimize both the penetration of the coolant jet and the mixing of the jet with the crossflow. In order to determine the optimum cooling hole design for the turbine, it is desirable to have a clear understanding of the film-cooling flow physics, and to have a predictive procedure that can accurately predict the aerodynamics and the heat transfer along the airfoil surfaces. In view of the turbulent nature of the flow, and the important role of the large scale unsteadiness, the development of predictive procedures need special attention. This paper provides a perspective of the available computational methods for film cooling and the accuracy levels expected with these methods. Film cooling is employed on airfoil surfaces and the endwalls including the hub, the blade tip and shroud. In general, several rows of cooling holes are employed. However, the basic configuration of interest that has been the focus for film cooling predictions is that of a single cold jet injected at an angle to the crossflow. The resulting flowfield is characterized by an array of large scale vortical structures [Fric and Roshko, 1994]. At the higher blowing ratios (defined as BR=(V)coolant/(V)crossflow)) of the coolant jet these structures could include the kidney shaped counter-rotating vortex pair (CVP), the horse-shoe vortex, the shear-layer vortices and the wake vortices (see Fig. 1 for a cartoon). These structures are inherently unsteady and anisotropic. The CVP is generally the dominant structure, and both the shearing between the jet and the crossflow and the vorticity issuing from the jet exit have been attributed to be the source of the CVP [Kelso et. al., 1996]. Upstream of the jet, due to the adverse pressure gradients, a horse-shoe vortex system is formed, which wraps around the base of the jet travelling downstream with vorticity counter to the CVP [Andreopoulos, 1985]. Shear layer vortices on the leeward and windward edges of the jet have also been observed, and have been attributed to Kelvin-Helmholtz instabilities. Periodic shedding of

Int. Symp. on Heat Transfer in Gas Turbine Systems 9 14 August, 2009, Antalya, Turkey

wake vortices occur when the jet blowing ratio is greater than 1. Considerable effort has gone into determining the origin of the wake vortices, and there is now experimental [Kelso et. al., 1996] and computational evidence [Tyagi and Acharya, 2005] that the wake vortices are initiated by the entrainment of the crossflow boundary layer into the wake, and the upward reorientation of the entrained flow into the wake structures. These and other studies provide unambiguous evidence of the importance of the coherent structures and their dynamics in the near field of the injected jet. These structures are clearly anisotropic as demonstrated computationally [Muldoon and Acharya, 2006] and experimentally [Kaszeta and Simon, 1999]. Clearly, any predictive model must embody the physics of the coherent structures to accurately predict the near-field jet behavior.

Figure 1: Schematic of the flow field of a jet in crossflow [Fric and Roshko. 1994)

While large scale structures play an important role in the film cooling dynamics, turbulence is critical in controlling the mixing behavior. Turbulence is characterized by a spectrum of scales, as shown in the cartoon in Fig. 2, with the more energetic lower wavenumber [larger length and time scales) eddies cascading energy down to the larger wavenumber, smaller scales. The challenge of any turbulence calculation is the ability to model the behavior of the entire spectrum of scales. The larger scales are anisotropic in nature, and are not well-represented by universal models. In view of this, accurate prediction of film cooling flows and heat transfer have been difficult to obtain. Most calculations have been reported by solving the Reynolds-Averaged Navier-Stokes Equations (RANS) that require a turbulence model. With this approach, the model must represent the effect of the fluctuations over the entire range of scales and, to date, such models have not been developed. While RANS calculations have typically been performed to obtain steady solutions, it is possible to obtain unsteady-RANS (URANS) results to resolve large-time scale fluctuations arising from timedependent boundary conditions or geometrical aspects of the flow (e.g., vortex shedding behind a cylinder). Nevertheless, both RANS and URANS can not resolve any part of the turbulent spectrum which must therefore be modeled. On the other hand, recent advances in processor speed and parallel computing have made very computer-intensive calculations possible, and it is now feasible to solve the unsteady Navier-Stokes equations with sufficiently fine mesh-spacings and time-steps to resolve the necessary spatial and temporal scales in the flow. In addition to resolving the flow scales, the numerical schemes have to be sufficiently accurate (low truncation error), so that, the energy is not artificially dissipated. Such higher-order schemes with the spatial and temporal resolution down to the dissipative scales (Fig. 2) are called Direct Numerical Simulation (DNS). RANS and DNS represent the two boundaries of the calculation schemes for turbulent flows. In DNS all scales of the flow are simulated exactly, and therefore the resolution requirements make DNS expensive. As Reynolds number is increased, the flow scales become smaller, and the spatial and temporal resolution requirements increase, eventually becoming prohibitively expensive. Thus DNS is currently possible only at low Reynolds numbers (typically Ret<500, where Ret is based on the

Int. Symp. on Heat Transfer in Gas Turbine Systems 9 14 August, 2009, Antalya, Turkey

wall friction velocity ut). In RANS, only the time-averaged mean flow is solved for, and the effect of the temporal and spatial fluctuations are embodied into the turbulence model. Thus the accuracy of the solutions are, to a large extent, controlled by the validity of the turbulence model. Large Eddy Simulation (LES) bridges the two extremes of DNS and RANS. In LES, only the larger-scale nonisotropic motions (that lies in the inertial subrange, Fig. 4) are resolved, while the smaller scale isotropic motions (closer to the dissipative subrange) are modeled by a dynamic, eddy-viscosity type model [Ghosal and Moin, 1995]. The resolved scales depend on the grid size used, and the cut-off between the resolved and modeled scales can be shifted by refining or coarsening the grid. LES offers the potential of realistically simulating a large array of engineering flows since, with fairly modest resolution, the inherent dynamics of the coherent structures are correctly captured. In wall bounded flows such as those encountered in film cooling, the length scale of the energy carrying structures become smaller as the wall is approached, and resolution requirements increase toward the wall. To address this issue, a hybrid approach called Detached Eddy Simulation (DES) that combines unsteady RANS (RANS) near the wall with LES away from the wall is sometimes adopted.
Large scales (Boundary condition dependent) Forward Transfer Backscatter Energy containing scales Ine rtia lR

Small Scales (Probably more universal)

Log(E(k))

sca ling )

Dissipative Scales (Kolmogorov Scales) Log(k)

Figure 2: Cartoon showing the scales in a turbulent flow, and the energy spectrum The present paper attempts to examine the available predictive models reported on film cooling, and the ability of these models to reproduce the measured trends. GOVERNING EQUATIONS The governing equations represent the conservation of mass, momentum and energy which are expressed either in an instantaneous form (DNS), spatially-averaged form (LES), or temporallyaveraged form (RANS), but in all cases, can generally be written in dimensionless form as:
U j =0 x j U i U iU j p 1 2U i ij + = + + + fi t x j x i Re x 2 x j j 1 2 q j +Uj = + t x j Re Pr x 2 x j j

SPECTRAL PICTURE

ang e ( k

-5 /

V is co us

PHYSICAL SPACE PICTURE

Di ss ipa tio n

(1)

Int. Symp. on Heat Transfer in Gas Turbine Systems 9 14 August, 2009, Antalya, Turkey

where Uj and are the non-dimensional resolved velocity and temperature respectively, xj is the spatial coordinate, p is the non-dimensional pressure, and ij and qj represent the effect of the unresolved fluctuations in the flow on Uj and respectively. As the mesh resolution is increased a smaller portion of the turbulent spectrum is modeled and the contribution of ij and qj reduces. In the limit where all scales (inertial and dissipative) are resolved (DNS), no modeling is required and the modeled terms are zero. Where the majority of the inertial subrange is resolved, only the small scales are modeled (LES), and ij and qj only incorporate the effect of the small scale fluctuations on the resolved scales. In the limit, where all scales are modeled (RANS), Uj and represent the mean (averaged) quantities, and the modeled quantities ij and qj must include the effect of the entire spectrum of scales including the larger-scale anisotropic eddies that are typically difficult to model with an universal model. In most cases where RANS is the simulation methodology, the unsteady term is generally dropped since only the time-averaged behavior is predicted. However, from a mathematical perspective, the time-averaging is done over a time-period long enough to average out the turbulent fluctuations, and therefore if the unsteady term is retained in the RANS equations, only large-scale unsteadiness with time-periods greater than the averaging period can potentially be resolved. Such unsteadiness can arise from boundary conditions or the geometry and an unsteady-RANS (URANS) that uses accurate differencing schemes (e.g., second order scheme) can be used to resolve these low-frequency fluctuations. The two quantities of primary interest to the turbine designer are the film cooling effectiveness () and the heat transfer coefficient. (h). The cooling effectiveness defined as (=(Taw-T )/(Tc- T ) is a representation of the non-dimensional wall temperature with an adiabatic surface. The heat transfer coefficient is calculated with a constant wall temperature boundary condition. Therefore, in order to get the complete heat transfer information, the simulations require both isothermal and adiabatic wall boundary conditions. For unsteady computations, the solutions for both boundary conditions have to be obtained simultaneously. BRIEF LITERATURE REVIEW RANS Studies: The majority of the reported film cooling simulation studies are based on the timeaveraged RANS equations. Since the literature is quite extensive, only representative studies are presented here. In a series of papers, Walters and Leylek [1997] performed a number of computational studies using turbulence modeling with a commercially available flow solver (Fluent). The grids used are unstructured, tetrahedral meshes and the geometry and flow conditions (blowing ratios and density ratios) corresponded to those of Pietrzyk et al. [1989] and Sinha et al. [1991]. A two-equation turbulence model was used. The jetting effect in the coolant delivery tube due to the sharp transition from the plenum to the tube was described in detail in this paper. Beyond 10 hole diameters downstream of coolant injection, the surface adiabatic effectiveness is well predicted, but the blow off and reattachment effects closer to the hole at higher blowing ratios is not captured. In Walters and Leylek [2000], a two-layer turbulence model is used which successfully resolves some of the jet blow off and reattachment effects and brings the surface adiabatic effectiveness plots closer to the experimentally measured values in the near field, but still leaves much to be desired in this respect. McGovern and Leylek [2000] analyzed compound angle holes at higher blowing ratios (1.25 to 1.88) and at a density ratio of 1.6. No validation to experiments is provided for the compound angle holes. Hyams and Leylek [2000] compared the results of shaped film cooling geometries that included forward diffused holes, laterally diffused holes, cylindrical holes with shaped inlet regions, and cusp-shaped holes, and these results were compared to the plain cylindrical geometry. Validation to experimental data generated by Schmidt et al. [1996] for the

Int. Symp. on Heat Transfer in Gas Turbine Systems 9 14 August, 2009, Antalya, Turkey

surface adiabatic effectiveness is provided for the cylindrical hole for two blowing ratios and shows reasonable agreement. Details of the flow field and plots of the surface adiabatic effectiveness and heat transfer are shown. A discussion of induction lift and its effect on the design of shaped film cooling holes is provided. Brittingham and Leylek [2000] investigate shaped geometries combined with compound angles. Validation to experimental data generated by Schmidt et al. [1996] for the surface adiabatic effectiveness is provided for the cylindrical hole and forward diffused hole and shows reasonable agreement. In a more recent study, Jones et al. [2005] investigate the suitability of turbulence modeling (specifically, a four equation model) for film cooling simulation. The turbulence model is first validated with respect to flow field via comparisons to experimental data. The effects of a variable Prandtl number are investigated in terms of agreement with experimental jet temperature data and surface temperature data. The console geometry is investigated in a computational study by Azzi and Jubran [2007] utilizing a two-layer, anisotropic RANS turbulence model. Good agreement is shown with the laterally averaged surface effectiveness data and the spanwise surface effectiveness data by Sinha et al. [1991] for the cylindrical hole. Little in-depth analysis of the flow fields is performed, but a tendency of the jet effused from the console hole to penetrate the crossflow excessively where the jets from two adjacent consoles meet is shown. No comparisons to data from the experiments of Sargison et al. [2002] are included. This is possibly due to the differences in geometry (most notably length to diameter ratio of the holes) and operating point (Azzi and Jubran, 2007 set the blowing ratio, while Sargison et al., 2002 set the ideal momentum flux ratio). LES Studies: There have been a number of studies that have utilized LES for film cooling. Tyagi and Acharya [2003] performed LES of film cooling with a discretization scheme that is fourth-order accurate in space and third-order accurate in time and uses a dynamic mixed model for the computation of subgrid stresses. Inlet conditions from a RANS calculation are used at the bottom of the flow domain in the plenum, and a random perturbation generator is used to generate the turbulence for the delivery tube inlet. Predictions show the evolution of the flow structures and the role they play in the near-wall surface effectiveness. Validations of velocity profiles as well as centerline surface adiabatic effectiveness profiles are presented and agree well with published results of Sinha et al. [1991] and Lavrich and Chaiappetta [1990]. More details on these predicted results are provided later. Iourokina and Lele [2005] reported a LES study via a two-domain approach for film cooling. A compressible, fourth-order code based in Cartesian coordinates is used in the region outside the delivery tube, while a low Mach number, second-order accurate code based in cylindrical coordinates is used in the delivery tube and cylindrical feeding plenum chamber. The dynamic Smagorinsky model is used in the subgrid scale modeling for both codes. The delivery tube, plenum chamber, and crossflow are all independent domains linked together by overlapping grids. The solutions at each time step are forced (by interpolation from one grid to the other and vice versa) to match in the intersecting volumes. In Iourokina and Lele [2006a] they have examined the role of freestream turbulence via a spatially developing turbulent boundary layer approach as presented by Lund et al. [1998]. The experimental setup of Pietrzyk et al. [1989] is simulated in order to provide validation of the flow field. A density ratio of 0.95 and a blowing ratio of 0.7 is used to allow comparisons to both the flow field data of Pietrzyk et al. [1989] and the heat transfer data with high freestream turbulence of Bons et al. [1996]. Details of the flow field are discussed, but no significant surface effectiveness or heat transfer data is presented. The crossflow domain is limited to a very small region outside of the cooling hole, only 5 hole diameters long in the flow

Int. Symp. on Heat Transfer in Gas Turbine Systems 9 14 August, 2009, Antalya, Turkey

direction total. In Iourokina and Lele [2006b] a more detailed examination of the flow field characteristics was performed with an attempt at validating the centerline surface heat transfer data of Sinha et al. [1991] at a lower blowing ratio where there was no significant jet detachment. Peet and Lele [2008] further examine the flow field predictions through extensive validation of the flow field (including both first and second order statistics) with experiments done by Pietrzyk et al. [1990]. Comparisons of film cooling effectiveness to the experiments of Sinha et al. [1991] are made with good agreement in the near field (up to three hole diameters downstream of the hole). Over two million cells are used in the small crossflow region simulated indicating a highly-resolved mesh. It should be noted that because the filter width is dependent on the grid spacing, as the mesh spacing gets finer and finer, smaller scales of turbulence are being resolved and when the mesh gets fine enough and approach the Kolmogorov scales (the scales at which the fluctuations are dissipated into heat), the simulation is close to being a DNS. The LES study of Guo et al. [2006] simulates two simple geometries- a normal jet in crossflow and a 30 inclined jet in crossflow. Both jets are issued through a plate 12 hole diameters thick with a feeding plenum chamber at the bottom. A spatially developing turbulent boundary layer simulation is carried out independently from the jet simulation and is used to generate turbulent inlet boundary conditions via the method of Lund et al. [1998]. The flow fields are examined in detail, but not compared to any other studies. Renze et al. [2006, 2007] utilize the same code and spatially developing turbulent inlet boundary conditions, and simulated both an inclined hole and a shaped hole (laterally and forward diffused) at a single blowing ratio (0.43) and density ratio (1.53, consistent with CO2 injected into air). Comparisons of the mean velocity and second order statistics of velocity are made to an experimental study and show good agreement. Results of surface adiabatic effectiveness are plotted using the mass transfer analogy and plotting the mixture fraction of CO2. In a more recent study, Renze et al. [2008a] present LES results for a cylindrical hole at blowing ratios from 0.15 up to 0.5 with density ratios of unity and 1.53. Validation to an experimental PIV study is provided and shows good agreement for first and second order statistics of velocity. The impact of varying velocity ratio and density ratio are investigated in some detail, with a focus on understanding the basic physics of the jet in crossflow. A comparison to the experimental work of Sinha et al. [1991] is made, but for a different blowing ratio and different geometry than the experiments. In Renze et al. [2008b] a comparison of surface adiabatic effectiveness to the data of Sinha et al. [1991] for a blowing ratio of 0.5 is presented. However, there are differences in the configurations with a longer coolant delivery tube being used in the predictions. Leedom and Acharya [2009a] in a very recent study have performed LES for cylindrical holes with L/D ratios of 1.75 and 3.5 using a dynamic Smagorinsky model and no freestream turbulence. They have performed these simulations for a blowing ratio of 0.5 and 1, and density ratio of 2 and compared with the flow measurements of Pietzryk et al. [1989] and heat transfer measurements of Sinha et al. [1991]. Agreement with flow data is excellent although the comparison with the cooling effectiveness results is only modestly good. Representative results are presented later. As an extension of this study, Leedom and Acharya [2009b] undertook LES of laterally-diffused shaped holes and console-shaped holes. Rozati et al. [2007a, 2007b], in recent years, have also studied LES of film cooling arising from the leading edge of an airfoil. They have shown these results to be in reasonable agreement with published data.

Int. Symp. on Heat Transfer in Gas Turbine Systems 9 14 August, 2009, Antalya, Turkey

DNS Studies: DNS studies of coolant jets are limited due, in part, to the resolution requirements at the Reynolds numbers of interest. Muppidi and Mahesh [2005] study a normal cylindrical jet in crossflow using DNS. Important flow features are investigated including the horseshoe vortex, hovering vortex, and the jet exit profile. Effects of crossflow profile and velocity ratio (equal to blowing ratio in this case at density ratio unity) on these features are presented. A scaling law for the jet trajectory is presented. Muppidi and Mahesh [2007] also studied the DNS of a normal, round jet. More discussion of jet exit profile (first and second order statistics) and condition are presented. In Muppidi and Mahesh [2008] simulations are presented of both the flow field and the transport of a scalar in a normal, round jet. The jet is at a very high blowing ratio (5.7) making the results of the study of limited usefulness to film cooling applications. Muldoon and Acharya [2006] presented DNS of an inclined film cooling jet and extracted budgets of the turbulence kinetic energy and dissipation rates so that closure models can be examined. They demonstrated that the (k2/) scaling for the eddy viscosity is the major culprit for the inaccuracy in the k- calculations. RESULTS AND DISCUSSIONS In this section we will present representative results from RANS, LES and DNS studies with a focus on determining the predictive capabilities of each method. RANS Results: Figure 3 shows the predictions of cooling effectiveness from two-equation models compared against the experimental data of Sinha et al. [1991] at a blowing ratio (M=(V)j/(V)cf) of 1. Discrepancies observed are quite significant and none of the models reproduce the observed separation and reattachment downstream of the hole. For the two predictions shown, the cooling effectiveness significantly overpredicts the measured values in the near-field of the coolant hole. Fig. 4 shows the cooling effectiveness behavior at a lower blowing ratio of M=0.5 (from Lakehal, 2002]. The typical two-equation model (denoted by the solid line and referred to as TLV) overpredicts the centerline effectiveness (Fig. 4a) and under-predicts the lateral average (Fig. 4b) which is also related to a prediction of smaller lateral spreading or coverage. This under-prediction of the lateral coverage is linked to the significantly lower levels of lateral stress predicted by these models. At higher blowing ratios, greater vertical jet penetration is predicted by the turbulence models leading to a larger separation and under-predicted values of effectiveness close to the jet exit. This is clearly shown in Fig. 5 at a M of 1.

Figure 3: Centerline cooling effectiveness predictions with two-equation models (M=1)

Int. Symp. on Heat Transfer in Gas Turbine Systems 9 14 August, 2009, Antalya, Turkey

In order to improve the predictions of the lateral spreading of the coolant jet (and mixing), several investigators have proposed anisotropic corrections to the expression for eddy viscosity. Bergeles et al. [1978] proposed multiplying the eddy viscosity appearing in the cross Reynolds stresses and the heat fluxes by a linearly varying factor obtained by a fit to the experimental pipe-flow data. More recently, Lakehal and coworkers, in a series of papers [Lakehal, 2002; Aziz and Lakehal, 2002; Lakehal et al. 1998], have used anisotropic corrections for eddy diffusivity and secondary stresses. In Lakehal [2002], near-wall DNS channel and boundary layer data have been used to parametrize the wall-normal velocity fluctuations (v2), and the wall-lateral velocity fluctuations (w2) is expressed as w2 / k = f (v 2 / k ) where f is the anisotropy factor that is provided empirically. The

t 3 = ( w2 ) l where l is a specified eddy diffusivity are then provided as t1 = t 2 = (v 2 ) l ; length scale. By this process, the model (TVLA in Fig. 4) uses different eddy diffusivity in the lateral and normal directions. Figure 4(a) and 4(b) show that both the centerline and spanwise distribution of effectiveness are improved with this process. If the Prandtl number is further parameterized with DNS data, further improvements are obtained with the model predictions (TLVA-Pr). Further support of improvements obtained with Prandtl number functionalization is presented by Jones et al. [2005] who demonstrated that a Ret-dependent turbulent Prandlt number produced improvements in the predictions of film cooling effectiveness.

1/ 2

1/ 2

Figure 4: Distribution of centerline and spanwise cooling effectiveness (M=0.5) with a two-layer k-e (TLV) model, with anisotropic corrections to the eddy viscosity (TLVA) and the turbulent Prandtl number (TLVAPr), from Lakehal [2002]

We conclude from the above discussions that the traditional two-equation turbulence models do not perform well for film cooling applications. The high-Re version of the model over-predicts the coolant jet kinetic energy; the low-Re version models while correctly predicting the peak levels of the kinetic energy in the jet, under-predict the turbulence levels in the near-wall and wake regions. In all cases the lateral turbulent stresses are under-predicted by the models and are likely responsible for the reduced lateral spreading of the coolant jet observed in the two-equation model predictions.

Int. Symp. on Heat Transfer in Gas Turbine Systems 9 14 August, 2009, Antalya, Turkey

Fig. 5: Comparison of RANS predictions of film cooling effectiveness with measurements, from York and Leylek, 1999

LES Results: To demonstrate the predictive capability of LES, a few representative comparisons with corresponding time-average experimental data are presented here. For a 35-degree inclined coolant injection, Tyagi and Acharya [2003] presented the time-averaged computational results using LES and compared the flow field data with the velocity measurements of Lavrich and Chiapetta [1990]. In Fig. 6, comparison of the streamwise (U), and vertical (V) components of velocity field are shown along the centerline (Z/D = 0.0), while the spanwise (W) component of velocity field is shown at Z/D = 0.5 at three streamwise stations (X/D = 0, 5 and 10). Good agreement between the experimental data and the time-averaged computational results demonstrate the cost-effective predictive capabilities of LES. Predictions of the centerline film cooling effectiveness at a blowing ratio of 0.5 are compared against the experimental data of Sinha et al. [1991] in Fig. 7. In these simulations as well as experiments, the coolant delivery ducts are short and equal to 1.75 times the duct diameter. The excellent agreement between the cooling effectiveness predictions and the experimental data provides another strong argument in favor of the predictive capabilities and accuracy of LES for such unsteady flows. In the more recent study by Leedom and Acharya [2009a], a very detailed comparison of LES predictions of a cylindrical film cooling hole has been made with the flow data of Pietzryk et al. [1989]. The flow data consists of mean velocity and turbulence in all three directions. In the calculations, freestream turbulence was not included, and therefore all turbulence generated was due to the flow development and jet-crossflow interactions. Figure 8 shows the mean velocity predictions (U and V) compared with data for one blowing ratio (M=0.5), while Figure 9 shows the corresponding turbulence predictions for urms and uv [Leedom, 2009]. The predictions and data shown appear to agree very well with each other. While not shown here, the comparison of the predictions and measurements for all three velocity components and all components of the turbulence stresses at different blowing ratios and density ratios are excellent [Leedom, 2009] and are a strong testament to the ability of LES to accurately predict such flows. A key observation of significant relevance here is how well all the non-isotropic turbulent stress components are predicted. It should be noted that RANS predictions do not correctly predict the anisotropic stress behavior. Since the predicted and measured stresses are in agreement with each other, and since no

Int. Symp. on Heat Transfer in Gas Turbine Systems 9 14 August, 2009, Antalya, Turkey

freestream turbulence was specified in the calculations, it indicates that the majority of turbulence is generated by the jet-crossflow interaction.

a) U

b) V

c) W
Figure 6: Comparison of the numerical predictions with the experimental data along the centerline for M = 1.0 [Tyagi and Acharya, 2003].

Figure 7: Film cooling effectiveness along the centerline of the computational domain for M = 0.5 [Tyagi and Acharya, 2003].

Int. Symp. on Heat Transfer in Gas Turbine Systems 9 14 August, 2009, Antalya, Turkey

Figure 8 : Mean streamwise velocity (U) and vertical velocity (V) at several axial locations (X/D) downstream of the cylindrical coolant hole. Symbols are experimental data of Pietzryk et al. [1989] and dots are LES prediction from Leedom [2009]. Blowing ratio is 0.5, density ratio is 2.

Figure 9: Turbulent stresses (urms and uv) at several axial locations (X/D) downstream of the cylindrical coolant hole. Symbols are experimental data of Pietzryk et al. [1989] and dots are LES prediction from Leedom [2009]. Blowing ratio is 0.5, density ratio is 2.

Int. Symp. on Heat Transfer in Gas Turbine Systems 9 14 August, 2009, Antalya, Turkey

Figure 10: Predictions of film cooling for cylindrical hole, blowing ratio of 1. Comparison of experimental data [circular symbols, Sinha and Bogard, 1991], LES predictions (from Leedom, 2009]

Figure 10 shows the cooling effectiveness predictions for a cylindrical hole using LES and these predictions are compared with the published RANS data and measurements of Sinha et al. [1991] for a blowing ratio of 1. Clearly, none of the RANS predictions are able to capture the measured trend of separation and reattachment, while LES results show the same behavior as the measured trend. In the near-field of the coolant jet, till about x/D of 7 or so, the LES predictions are clearly more superior to any of the RANS predictions. LES results at a blowing ratio of 0.5 were presented earlier in Fig. 7 and show excellent agreement with the data of Sinha et al. [1991]. These reasonably good agreement with data at two blowing ratios provides a measure of confidence that LES results can have broader applicability for film cooling predictions. As stated earlier, JICF serves as the canonical problem for the film cooling flows and thus, the generation and evolutionary mechanisms of various coherent structures is usually explained using flow features such as counter rotating vortex pair (CVP), wake vortices, horseshoe vortices and roller vortices. Tyagi and Acharya [2003] identified a hairpin shaped coherent structure as the main building block for the large scale flow field (Fig. 11). The morphological details of the hairpin coherent structures along with complimentary orthogonal projected views of the induced flow in the near vicinity explained all of the above-mentioned prototypical vortices in JICF in the unified fashion. The streamwise CVP is associated with the horizontal legs of the hairpin structure. The upright legs of the hairpin structures induce wall normal vorticity and convect downstream while entraining the crossflow fluid. Some of the entrained fluid is reoriented along these upright legs forming a structure similar to the wake vortices. In the projected view normal to the center plane of the hairpin structure, induced flow associated with the head of the hairpin (U-shaped bend) is similar to the roller vortices. Thus, these hairpin structures can explain previously considered different vortices as no more than the orthogonal projections on different view-planes. Hairpin vortices are believed to be the building block of turbulent boundary layers. The region near the surface is a complex environment dominated by the presence of a myriad of vortices that are believed to be predominantly of the hairpin type. Such vortices constitute moving lagrangian disturbances which carry concentrated vorticity in the core region that diffuses progressively outward with time. These vortices are embedded in a highly sheared background flow near the surface; over time they can be expected to distort into complex shapes, as well as to interact with one another, resulting in the evolution of complicated vorticity topologies. Smith et al. [1991] studied the evolution of such hairpin vortices that are generated by impulsively injecting fluid into a subcritical laminar boundary layer (Fig.

Int. Symp. on Heat Transfer in Gas Turbine Systems 9 14 August, 2009, Antalya, Turkey

12). Clearly, one can expect that similar hairpin structures should be present in the jet-in-crossflow situation with significantly increased complexity of both the background shear flow as well as the topology of lagrangian disturbance issuing out of the jet. Zhou et al [1999] studied the evolution of a single hairpin vortex in the mean turbulent field of a low Reynolds number channel flow using direct numerical simulation. The initial flow structure was a viscous hairpin vortex structure extracted from the full two-point correlation tensor of a low Reynolds number channel flow database using the linear stochastic estimation (LSE) procedure. This initial vortex structure in a clean turbulent mean flow environment is studied to visualize the complex evolution and subsequent auto-generation of new hairpin vortices. They observed that new hairpins generate downstream of the primary hairpin above a threshold strength of this structure, thereby forming, together with the upstream hairpins, a coherent packet of hairpins that propagate coherently. These vortices were observed to pass low-speed fluid from the downstream vortex to its upstream neighbor and so on over several hairpin vortices to form near-wall low-speed streaks of length significantly longer than the streamwise lengthscale of any single hairpin vortex. Since the wake region of jet-in-crossflow is populated periodically with such hairpin vortices that can autogenerate further, this mechanism is very critical in understanding the dynamics and influence of these vortices on the skin-friction and near-wall heat transfer for the film-cooling applications.

Figure 11: Details of the hairpin coherent structure and the orthogonal projection views of induced flow in the vicinity [Tyagi and Acharya, 2003].

(a) Plan view

(b) Side view

Figure 12. Picture of dye-marked single hairpin vortex generated using controlled injection through a narrow streamwise slot into a subcritical laminar boundary layer [Smith et al., 1991].

Generation of these hairpin vortices depends crucially on several flow parameters and it can rationalize the view of numerous parametric studies. The deformation and strength of issuing

Int. Symp. on Heat Transfer in Gas Turbine Systems 9 14 August, 2009, Antalya, Turkey

crossflow jet vortex ring primarily depends on the jet Reynolds number, injection angle, jet-tocrossflow velocity ratio, upstream crossflow-to-jet boundary layer thickness ratio and hole geometry [Haven and Kurosaka, 1997, Andreopoulos, 1985]. The deformed jet vortex ring can lead to observed hairpin vortices for inclined circular jets in crossflow. The coherent structures associated with Lagrangian disturbances of jet and crossflow interactions issue out of the hole as the fluid envelopes around the hairpin-shaped loci of local pressure minima. The streamwise spacing between consecutive coherent structures and the Strouhal frequency of the wake vortex shedding process are related through their convection speed. Hairpin structures evolve while convecting downstream in the wake region and control the entrainment and mixing of the crossflow fluid with the injected coolant fluid. Entrained unmixed crossflow fluid creates local pockets in the wake region that lead to the formation of hot spots on the film cooled surface. For a 35-degree simple injection, a time sequence of coherent structures overlaid on the surface temperature distribution reveals the metamorphosis of these hairpins during their life cycle as well as their influence on the instantaneous surface wall temperature (Fig. 13). The evolution of vortex structures is also shown in this figure with the corresponding structures at each time instance denoted by the same letter symbol. The experimental visualizations and measurements reported in the literature [Smith et al, 1991] support the hairpin structures argument outlined by Tyagi and Acharya [2003] and shown in Fig. 16. However, most of the studies in the literature deal with the vertical jet in crossflow scenario. Marzouk and Ghoniem [2007], Yuan and Street [1998] and Yuan et al. [1999] are examples of these studies. In Yuan et al. [1999], LES of round transverse jet in crossflow is performed, and hanging vortices are identified as the origin for ubiquitous far-field CVP . Note that this view is in contrast with the origin of CVP in the issuing jet vorticity. In Fig. 13, the important role of the coherent structures in dictating the jet evolution and the wall temperature distribution is clearly identified. Much of the LES discussion so far has been focused on cylindrical hole configurations. Leedom [2009] has undertaken LES simulations for fan-shaped and console-shaped holes. It is observed that for fan shaped holes the hair pin vortex structure is not as well defined as for the cylindrical holes; rather a collection of hair-pin vortices are observed in the flow field. This is shown in Fig. 17. As seen from the above studies, LES offers the potential of doing physically realistic engineering computations with moderate computational effort. With advances in computing power, in terms of processor speed, network speed and networked-cluster computers, LES for film cooling can now be accomplished on a large cluster in a time frame of the order of 1-2 days. Modeling of Surface Roughness: In recent years, the use of dirty fuels with increased particulate concentrations leads to enhanced surface deposition and erosion on airfoil surfaces and can produce rough surfaces that can alter film cooling characteristics significantly. Both large scale and small scale roughness has been observed as a result of deposition and erosion. The study of film cooled rough surfaces has been predominantly experimental till now. For computational investigations, the resolution and meshing requirements of surface asperities is a major challenge. Kalghatgi and Acharya [2009] have recently used LES with a roughness model to study the effect of roughness on the film cooling. Roughness is decomposed to have a macro- and a micro- component based on the ability of the mesh to resolve the peak-to-peak magnitude of the roughness elements. Macro roughness is modeled using an Immersed Boundary Method (IBM), and a roughness element model is used for micro roughness over immersed surfaces.

Int. Symp. on Heat Transfer in Gas Turbine Systems 9 14 August, 2009, Antalya, Turkey

Figure 13. Unsteady dynamics of coherent structures and their influence on wall temperature at different time instants. Arrows are tracking a specific structure to highlight the lifecycle of a hairpin coherent structure [Tyagi and Acharya, 2003]

Figure 14. Unsteady coherent structures for a fan shaped hole. Iso-surfaces of the pressure Laplacian at one time instance. From Leedom [2009]

In the roughness element model, used to model micro-scale roughness, the effect of additional drag due to roughness is modeled using additional body force term corresponding to a regular sandgrain type roughness elements (conical element is considered in study by Kalghatgi and Acharya, 2009]. The additional drag term is given as, (4)

Int. Symp. on Heat Transfer in Gas Turbine Systems 9 14 August, 2009, Antalya, Turkey

where A and V are frontal area and volume of the immersed part of the cone in computational cell. A Cd = 0.2 is assumed throughout the calculations. This drag is imposed for the cells that contain the roughness element. The dimension and distribution of these roughness elements is determined based on statistical roughness data due to Bons et al. [2001]. Key roughness parameters include a measure of the roughness height (k), cone diameter (d), and inter-cone spacing (s). These parameters can be related to typical sandgrain roughness parameters k/ (where is the momentum thickness), Ra (average roughness height), and s (roughness density parameter). Macro-roughness elements that can be resolved on the grid (implying that they span several cells) have been incorporated using the IBM by Kalghatgi and Acharya [2009]. A brief description of the IBM was given in part I of the paper, but it primarily consists of tracking the grid nodes adjacent to the macro-roughness and using a second-order interpolation function to determine the values at these nodes (that are not solved for based on momentum and energy equations) between the no-slip surfaces and the solved fluid points. More details on this method are provided by Tyagi and Acharya [2003].

Figure 15: Centerline film cooling effectiveness distribution for blowing ratio of 0.5, cylindrical hole, and different roughness heights (k/ ranging from 0 to 6)

Figure 15 shows the effect of different roughness heights on the centerline film cooling effectiveness ( or eta). The roughness height is represented by k/ (denoted as k/theta in the figure). In all cases the roughness elements are applied from upstream of the film cooling injection. It is clear that roughness degrades the cooling effectiveness at this blowing ratio. At low roughness values the cooling effectiveness profile shows the same shape as the smooth case. At higher values, because of the roughness induced thickening of the approach boundary layer, significant reductions in cooling is obtained. Control of Film Cooling Flows: Since DNS and LES provide spatio-temporal evolution of the coolant flows, it is possible to examine how active perturbation of such flows influences the film cooling process. In a recent study, Muldoon and Acharya (2009) examined the effect of coolant-jet

Int. Symp. on Heat Transfer in Gas Turbine Systems 9 14 August, 2009, Antalya, Turkey

perturbation on the cooling effectiveness behavior. Results are shown in Fig. 16 which demonstrate that with active perturbation there is the potential of increased effectiveness in the near field of coolant injection. This enhancement in effectiveness (both centerline and spanwise distributions are shown in Fig. 16) is observed to a greater extent for blowing ratios in the range of 1-2 where the coolant jet blows off the surface before reattaching, and the perturbations provide a mechanism for a lesser degree of separation and an earlier reattachment point.
M=1.5, St=0.000, DC M M=.75, correct .5 DNS and LES codes can also be coupled with optimization algorithms1to determine the St=0.000, DCvalues .45 .45 Slc: X=2 of the control parameters that provide the highest cooling effectiveness. Muldoon [2008] and Slc: X=6 Nikitopoulos et al. [2007] have used a coupled DNS-optimization algorithm to.3 examine the optimum Slc: X=10 .3 Slc: pulsing parameters for achieving maximum cooling effectiveness.X=14 approaches are currently being Such .15 .15 explored by the author.

0 -1.5

-1

-.5

M=.75, St=0.080, DC .5 M=1.5, St=0.000, DC 1


.45 .45

.5

0 -1.5

-1

-.5

M=.75, St=0.320, DC .5 M M=.75, St=0.000, DC .5


.45 .45

.5

.8

.6

M M M M

=1.5 = .75 = .75, St = 0.080, DC .5 = .75, St = 0.320, DC .5

.3 .3

Slc: X=2 Slc: X=6 Slc: X=10 Slc: X=14

.4

.15 .15 0 -1.5 0 -1.5

.2

-1 -1

-.5 -.5

M=.75, St=0.080, DC .5
0 2 4 6

0 z z

.5 .5

1 1

.3 .3 .15 .15 0 -1.5 0 -1.5 -1 -1

-.5 -.5

M=.75, St=0.320, DC .5
.45

0 z z

.5 .5

1 1

10

12

14

.45

Fig. 16: Centerline (left) and spanwise (right) distributions .3 cooling effectiveness. Pulsed cases (St=0.08 and of .3 0.32) have peak M of 1.5 and duty cycle (DC) of 0.5 leading to an average M of 0.75. Pulsed cases have half the .15 .15 coolant flow rate of the M=1.5 unpulsed case.

0 -1.5 -1 -.5

CONCLUSION

.5

0 -1.5

-1

-.5

.5

Strategies for the simulation of film cooling have been discussed and their predictive ability in terms of agreement with published data is presented. Specifically solutions resulting from RANS, LES, and DNS are outlined. The following major conclusions are reached: Calculations for film cooling must include the plenum since jet-exit conditions are difficult to specify, and appropriate boundary conditions must be provided at the crossflow inlet. For RANS, these boundary conditions represent the mean values of velocity and turbulence quantities, while for LES and DNS spatio-temporal values must be provided. RANS calculations using two-equation turbulence models exhibit several deficiencies including: o The vertical coolant jet penetration is typically over-predicted while the spreading in the lateral direction is lower than that measured. This leads to cooling effectiveness predictions that are typically lower immediately downstream of the hole and higher farther downstream compared to measurements. Also the lateral distribution of the cooling effectiveness is smaller in the predictions. o Realizability corrections that limit eddy viscosity magnitudes tend to improve the predictions. o Anisotropy corrections where the eddy viscosity in the wall-normal and spanwise directions are different improve the predictions of the cooling effectiveness considerably.

Int. Symp. on Heat Transfer in Gas Turbine Systems 9 14 August, 2009, Antalya, Turkey

o Modifications to the turbulent Prandtl number also move the predictions of the cooling effectiveness in the right direction. LES calculations resolve the energy carrying structures and only model the small-scale fluctuations which are generally isotropic in nature. In general, LES results require greater computational effort but show improvements over RANS, and the following observations are noted: o LES predictions of the mean velocities, turbulence and cooling effectiveness are in excellent agreement with measured data using a dynamic model for the Smagorinsky coefficient. No ad hoc corrections are needed for obtaining good agreement. o LES calculations provide saptio-temporal data and the associated evolution of the flow structures across a spectrum of scales. Thus they essentially capture the rich multi-scale physics of the fluid flow. DNS calculations require no modeling and resolve all scales, but are extremely computer intensive particularly at higher Reynolds numbers. As such, they are useful for: o Understanding, in detail, the flow physics across the spectrum of scales o Using the calculated budgets of the turbulence quantities to guide turbulence model developments o Using simulations combined with optimization algorithms for flow control studies (can also be done with LES). ACKNOWLEDGEMENTS The material presented in this lecture is compiled from various sources including those in the published literature and from work done by students in the authors research group. All these sources are gratefully acknowledged. In particular, the author would like to acknowledge the various contributions of his students whose work is represented in this paper including: Mayank Tyagi, David Leedom, Prasad Kalghatgi, Raymond Jones, and Frank Muldoon. The author is grateful for financial support and computational resources provided by various sponsors including: NASA, NSF, AFOSR, ONR, DOE and the Clean Power and Energy Research Consortium (CPERC).

Int. Symp. on Heat Transfer in Gas Turbine Systems 9 14 August, 2009, Antalya, Turkey

REFERENCES Andreopoulos, J. [1985], On the structure of jets in a cross-flow, J. Fluid Mech., Vol. 157, pp. 163197 Azzi, A. and B. A. Jubran. [2007], Numerical Modelling of Film Cooling from Converging Slot Heat Mass Transfer, Vol. 43, pp. 381-388. Brittingham, R. A. and J. H. Leylek. [2000], A Detailed Analysis of Film Cooling Physics: Part IV - Compound-Angle Injection With Shaped Holes. ASME Journal of Turbomachinery, Vol. 122 pp. 133-145. Bons, J. P., C. D. MacArthur, and R. B. Rivir [1996], The Effect of High Free-Stream Turbulence on Film Cooling Effectiveness, ASME J. of Turbomachinery, Vol. 118, pp. 814-824. Bons, J. P., Taylor, R.P., McClain, S.T. [2001], Many Faces of Turbine Surface Roughness. ASME Journal of Turbomachinery Vol. 123, pp. 231-242 Fric, T. F., and Roshko, A., [1994], Vortical structure in the wake of a transverse jet, J. Fluid Mech., Vol. 279, pp. 1-47 Ghosal, S. and Moin, P., [1995], The basic equations for the large eddy simulations of the turbulent flows in complex geometry, J. Comp. Phy., Vol. 118, pp.24-37 Guo, X., Schroder, W., and Meinke, M., [2006], Large-Eddy Simulations of Film Cooling Flows, Comput. Fluids, Vol. 35, pp. 587-606. Haven, B.A., and Kurosaka, M., [1997], Kidney and Anti-Kidney Vortices in Crossflow Jets, J. Fluid Mech., Vol. 352, pp. 27-64. Hyams, D. G. and J. H. Leylek. [2000], A Detailed Analysis of Film Cooling Physics: Part III Streamwise Inlection with Shaped Holes." ASME Journal of Turbomachinery, Vol. 122 pp. 122-132 Iourokina, I.V. and S. K. Lele. [2005], Towards Large Eddy Simulation of Film-Cooling Flows on a Model Turbine Blade Leading Edge. In Proceedings of AIAA 43rd Aerospace Sciences Meeting and Exhibit. Reno, Nevada. January 10-13, 2005. Iourokina, I. V. and S. K. Lele. [2006a], Large Eddy Simulation of Film Cooling Flow Above a Flat Plate from Inclined Cylindrical Holes. In Proceedings of ASME Joint U.S.-European Fluids Engineering Summer Meeting. Miami, Florida. July 17-20, 2006 Iourokina, I.V., and Lele, S.K. [2006b], Large Eddy Simulation of Film-Cooling Above the Flat Surface with a Large Plenum and Short Exit Holes, AIAA paper 2006-1102. Jones, R., S. Acharya, and A. Harvey. [2006], Improved Turbulence Modeling of Film Cooling Flow and Heat Transfer." In Modeling and Simulation of Turbulent Heat Transfer, by B. Sunden and M. Faghri, 113 Kalghatgi, P., and Acharya, S., [2009], LES of Film Cooling Over Rough Surfaces, Turbine-09, Antalya, Turkey, August 2009

Int. Symp. on Heat Transfer in Gas Turbine Systems 9 14 August, 2009, Antalya, Turkey

Kaszeta, R. W. and Simon, T. W. [1999], Measurement of Eddy Diffusivity of Momentum in Film Cooling Flows with Streamwise Injection, ASME J. of Turbomachinery, Vol. 122, pp 178 Kelso, R. M., Lim, T. T., and Perry, A. E., [1996], An experimental study of round jets in crossflow, J. Fluid Mech., Vol. 306, pp. 111-114 Lakehal, D., Theodoris, G.S., Rodi, W. [1998] Computation of film cooling of a flat plate by lateral injection from a row of holes, International Journal of Heat and Fluid Flow, Vol. 19 pp.418-30. Lakehal, D., [2002], Near-wall Modeling of Turbulent Convective Heat Transport in Film Cooling of Turbine Blades in Film Cooling of Turbine Blades With the Aid of Direct Numerical Simulation Data, ASME J. Turbomachinery, Vol. 124, 485-496. Lavrich, P.L., and Chiapetta, L.M., [1990], An Investigation of Jet in a Crosslow for Turbine Film Cooling Applications, United Technologies Research Center, UTRC Report No. 90-04. Leedom, D., and Acharya, S., [2009a], LES of Cylindrical Hole Film-Cooling: Plenum Effects, submitted for publication Leedom, D., and Acharya, S., [2009b], LES of shaped Holes, submitted for Publication, also ASME GT 2008-51009 Lund, T., Wu, X., and Squires, K., [1998], Generation of Turbulent Flow Data for Spatially Developing Boundary Layer Simulations, Journal of Computational Physics, Vol. 140, pp. 233-258 Marzouk, Y.M., and Ghoniem, A.F., [2007], Vorticity Structure and Evolution in a Transverse Jet, J. Fluid Mech., Vol. 575, pp.267-305. McGovern, K. T. and J. H. Leylek. [2000] "A Detailed Analysis of Film Cooling Physics: Part II Compound-Angle Injection with Cylindrical Holes." ASME Journal of Turbomachinery Vol. 122, pp. 113-121. Muldoon, F., and Acharya, S. [2009], Pulsed Film Cooling Flows, Intl. J. Heat Mass Transfer, Vol. 52, pp. 3118-3127 Muldoon, F. [2008], Control of a Simplified Unsteady Film-cooling Flow Using Gradient-based Optimization. AIAA Journal, Volume 46, Issue 10, pp 2443-2452. Muldoon, F., and Acharya, S., [2006], Budgets of the k-e model for Film Cooling and an Improved Damping Function Formulation, AIAA Journal, Vol. 44, No. 2, pp.3010-3031 Muppidi, S., and Mahesh, K., [2007], Direct Numerical Simulation of Round Turbulent Jets in Crossflow, J. Fluid Mech., Vol. 574, pp. 59-84. Nikitopoulos, D., Acharya, S., Muldoon, F. [2007], Pulsed Film Cooling Flows, AFOSR Contractor Meeting Report, 2007 Pietrzyk, J. R., D. G. Bogard, and M. E. Crawford, [1989], Hydrodynamic Measurements of Jets in Crossflow for Gas Turbine Film Cooling Applications, ASME Journal of Turbomachinery, Vol.

Int. Symp. on Heat Transfer in Gas Turbine Systems 9 14 August, 2009, Antalya, Turkey

111 pp. 139-145. Peet, Y. V. and S. K. Lele. [2008], Near Field of Film Cooling Jet Issued into a Flat Plate Boundary Layer: LES Study. In Proceedings of ASME Turbo Expo 2008. Berlin, Germany. June 9-13, 2008. Renze, P., Meinke, M., and Schroder, W., [2006], LES of Turbulent Mixing in Film Cooling Flows, Conf. on Turbulence and Interactions TI2006, May 29 June 02, 2006, Porquerolles, France. Renze, P., W. Schroder, and M. Meinke., [2007], Hole Shape Comparison for Film Cooling Flows Using Large-Eddy Simulations, In Proceedings of AIAA 45th Aerospace Sciences Meeting and Exhibit. Reno, Nevada. January 8-11, 2007. Renze, P., W. Schroder, and M. Meinke, [2008a], "Large-Eddy Simulation of Film Cooling Flows at Density Gradients." International Journal of Heat and Fluid Flow 29 (2008a): 18-34. Renze, P., W. Schrder, and M. Meinke, [2008b], Large-eddy Simulation of Film Cooling Flows with Variable Density Jets. Flow Turbulence and Combustion 80 (2008b): 119132. Rozati, A., Tafti, D. K., [2007], Large Eddy Simulation of Leading Edge Film Cooling Part-I: Computational Domain and Effect of Coolant Inlet Condition. ASME Paper No. GT200727689, Proceedings ASME Turbo Expo 2007: Power for Land, Sea and Air, May 14-17, 2007, Montreal, Canada Rozati, A., Tafti, D. K., [2007], Large Eddy Simulation of Leading Edge Film Cooling Part-II: Heat Transfer and Effect of Blowing Ratio. ASME Paper No. GT2007-27690, Proceedings ASME Turbo Expo 2007: Power for Land, Sea and Air, May 14-17, 2007, Montreal, Canada Sargison, J. E., S. M. Guo, M. L. G. Oldfield, G. D. Lock, and A. J. Rawlinson. [2002a], A Converging Slot-Hole Film-Cooling GeometryPart 1: Low-Speed Flat-Plate Heat Transfer and Loss." ASME Journal of Turbomachinery, Vol. 124, pp. 453-460. Schmidt, D., B. Sen, and D. Bogard. [1996]. Film Cooling with Compound Angle Holes: Adiabatic Effectiveness." ASME Journal of Turbomachinery, Vol. 118, pp. 800-806. Sinha, A. K., D. G. Bogard, and M. E. Crawford. [1991], Film Cooling Effectiveness Downstream of a Single Row of Holes with Variable Density Ratio. Transactions of the ASME, Vol. 113 pp. 442-449. Smith, C.R., Walker, J. D. A., Haidari, A. H. and Soburn, U., [1991], On the dynamics of near-wall turbulence, Phil. Trans.: Phys. Sci. Engg., Vol. 336, pp. 131-175 Tyagi, M., and Acharya, S., [2005], Large Eddy Simulation in Complex Geometries Using the Immersed Boundary Method, Intl. J. of Numerical Methods in Fluids, Vol. 48, pp. 691-722. Tyagi, M., and Acharya, S., [2003], Large Eddy Simulation of an Inclined Film Cooling Jet, ASME J. of Turbomachinery, Vol. 125, pp. 734-742. Zhou, J., Adrian, R. J., Balachandar, S. and Kendall, T.M., [1999], Mechanisms for generating coherent packets of hairpin vortices in channel flow, J. Fluid Mech., Vol. 387, pp. 353-396.

Int. Symp. on Heat Transfer in Gas Turbine Systems 9 14 August, 2009, Antalya, Turkey

Walters, D. K., and Leylek, J. H., [1997], A Systematic Computational Methodology Applied to a Three-Dimensional Film-Cooling Flowfield, ASME J. of Turbomachinery, vol. 119, pp. 777785 Walters, K., Leylek, J., [2000], A Detailed Analysis of Film Cooling Physics-Streamwise Injection with Cylindrical Holes, ASME J. of Turbomachinery, Vol. 122, pp. 102-112 York, W. D., and Leylek, J. H., [1999], Numerical Prediction of mainstream Pressure Gradient Effects in Film Cooling, ASME 99-GT-166 Yuan, L. L., and Street, R. L., [1998], Trajectory and Entrainment of a Round Jet in Crossflow, Phys. Fluids, Vol. 10, pp. 2323-2335 Yuan, L. L., Street, R. L., and Ferziger, J. H., [1999], Large-Eddy Simulations of a Round Jet in Crossflow, J. Fluid Mech., Vol. 379, pp. 71-104. Zhou, J., Adrian, R. J., Balachandar, S. and Kendall, T.M., [1999], Mechanisms for generating coherent packets of hairpin vortices in channel flow, J. Fluid Mech., Vol. 387, pp. 353-396.

Potrebbero piacerti anche