Sie sind auf Seite 1von 36

LITERATURE REVIEW

Introduction:
In this report I would like to summarize different research papers that where published on High experimental data while others is purely on experiments performed, since we need accurate and economical work, so the help of CFD and experiments is essential lift Airfoil Sections. Some research papers have done CFD analysis and validated the results with

1. The Aerodynamic Design of Multi-Element High Lift Systems For Transport Airplane, By C. P. Van Dam, Department Of Mechanical and Aeronautical Engineering University Of California,2002.
Summary: The purpose of this paper was to review recent developments in aerodynamics design and analysis methods for multi- element high lift systems on transport airplanes. Attention is also paid to the associated mechanical and cost problems since a multi element high lift system must be as simple and economical as possible, while meeting the aerodynamic performance levels. In the introduction the author signifies the importance of high lift systems. The Multi element high lift systems have significant impact on the cost of a typical jet transport because y y y y y They are time consuming to design and test Their flows, geometry and actuation and support systems are complex They are heavy Have a high part count They require intensive maintainance.

Conclusion: The aerodynamic design of multi-element high-lift systems for transport airplanes has evolved significantly from a largely empirical approach backed by extensive wind-tunnel and in-flight testing in the 1960s to a computational approach backed by limited wind-tunnel and flight tests today. However, the highly competitive and economically driven aviation market of today requires that aircraft manufacturers be able to provide customers with a high-quality product within the shortest amount of time at the lowest possible cost. To further reduce design cycle time and cost, the high-lift design and development process requires further improvements; particularly the computational three-dimensional aerodynamics methods that are being applied. Areas that need attention include:

Grid generation. Time required to progress from a CAD definition to a volume grid for the CFD analysis of a configuration. Especially for multi element wings with wing-mounted engine nacelles this step in the aerodynamic design and analysis process is still too time consuming. Turbulence modeling in RANS methods. Flows about multi-element high-lift systems are inherently separated and unsteady and wakes play a critical role in the aerodynamic interactions between the various high-lift elements. Most turbulence models in Reynolds- averaged Navier Stokes methods provide accurate predictions for attached boundary layer flows. However, the predictions tend to become less accurate for separated flows and wakes. Rumsey and Gatski have applied several turbulence models to multi-element airfoil flows and are continuing their efforts to make improvements to these models. Boundary-layer transition modeling in RANS methods. For most large jet-propelled transport airplanes with their highly swept wings, laminar flow and transition are hardly an issue at high speed conditions. At cruise the flow becomes turbulent at or shortly downstream of the leading edge as a result of attachment-line transition or cross flow instability. However, in the high-lift configuration surfaces such as slats and flaps experience much smaller Reynolds numbers and, hence, may support extended regions of laminar flow. This does affect the aerodynamic performance characteristics and, hence, a transition prediction capability is a necessary component of any computational aerodynamics method used in high lift design. Separation bubbles. More than 20 years ago, Dillner et al commented on the fact that much work remained to be done on the simulation of flows with separation or transition bubbles: While apparently a mere footnote to the overall high-lift problem, as long as wind tunnel tests continue to be conducted at low Reynolds numbers, the capability to predict the formation and effect of laminar separation bubbles remains an important, imperfectly developed, capability . Much progress has been made in this area but because these bubbles tend to be small and unsteady, RANS-based maximum lift predictions for flows governed by laminar separation bubbles remain problematic. Noting the importance of designing aerodynamically well performing, simpler and less costly high-lift systems, it is critical that high-lift system considerations be included at even the earliest stages of the design process. Presently, few tools are available that facilitate this type of concurrent approach at the conceptual and preliminary design stages.

2. High Lift Aerodynamics By A. M.O Smith McDonnell Douglas Corporation, Long Beach Calif June 1975.
Summary: The paper starts with Sir A. M. O Smith early days when he first began his work for the Douglas Aircraft Company in 1938. He then gives the idea of Aeronautical knowledge at time of his M. S. thesis project, to perform test on a powered Boundary-Layer-Control Model in the GALCIT Ten Foot wind tunnel. Transition as an explicit phenomenon in the development of a boundary layer was only vaguely recognized. Drag coefficient were measured at a series of tunnel speeds. Results where plotted on a log scale and then generally extrapolated by a straight-line extension to full scale. The curve usually had a downward slope with Reynolds number, but not always. Even when the slope was positive, the line would be extended the same way, the effect was dismissed as a Poor Reynolds No Extrapolation . By this we can see author was keen on research in aerodynamics. He wanted to know why things are changing, what is the physics behind all these reasons. He was against the idea of simply carrying out the experiments in wind tunnel and drawing results which has no reason. He says that the defects in knowledge were two Although what to do may have been known, the reason for doing it were dimly understood. Quantitative analysis of a flow could rarely be accomplished. Some Lift Limits: Just as ideal- cycle limits are useful in thermodynamics, so are the theoretical limits of lift useful in aerodynamics. Knowledge of those limits helps gives us a perspective as to where we are now and what may be attainable if we are willing to seek without compromise the maximum possible lift. CL= 2 Sin( + )/Cos [0< <45 ] CL= 4 Sin Sin( + ) [45 < <90 ] For symmetrical Joukowski airfoil CL= 2 (1+0.77t)Sin CLmax= 4 , =90 , =0

Conclusions: One could talk indefinitely about other special problems related to high lift but that wouldn t do, so he listed what he thinks are the 10 most important basic theoretical problems in High Lift Aerodynamics. 1) Very general calculation of three-dimensional laminar and turbulent flows. 2) Calculation of flows involving partial separation in the rear. 3) Practical calculation of flows involving forward separation bubbles. 4) Practical calculation of flows involving shock boundary- layer interaction. 5) Calculation of the viscous flow on airfoils and bodies of revolution of length c from about 0.95c Jo 1.05c. In that region of very rapid change and strong interaction, boundary layer equations do not apply. With our recently acquired ability to calculate well the forward 95%, our inability to solve the next 10% is a real irritant. 6) Further development of inverse methods. 7) Drag of multi element airfoil systems. Drag predictions have a relative error one order of magnitude greater than that of lift predictions. Since propulsive and acceleration information is crucial to aircraft design, that is a very important problem. 8) Practical calculation of merging boundary-layers, wall jets and wakes. Although methods have been developed, it is felt that considerable further development is needed. 9) The analysis of flows over swept wings on which a leading-edge vortex is developed. When the vortex develops, conventional calculations are about 100% in error. 10) Three-dimensional transonic calculations, particularly for arbitrary wings and wing-body combinations. In several of those problems, the first phase is two dimensional analyses, but as soon as that is accomplished the three-dimensional looms up as the next target.

3. Design and Analysis Of Low Reynolds Number Airfoil, By Nicholas . K . Borer. December 2002.
Summary: In this paper the author carries outs the analysis of the design of a low Reynolds number (1, 00,000 to 10, 00,000) airfoil to be used on Uninhabited Aerial Vehicles or UAV`s. The motivation for this study is the application of this airfoil or series of airfoils to a vehicle to be designed built and flown by members of Georgia Tech`s Design /built/Fly Team. Conclusion: The results are for incompressible, invicid flow which where compared against output from XFOIL for an invicid case and a viscous case for a Reynolds No of 10, 00,000. MATLAB Parameter CL Cm Cd CL Cm Cd Target 0.6000 * 0.000 0.600 0.000 0.000 Converged 0.6066 -0.1387 0.0008 0.6169 -0.0043 0.0008 Inviscid 0.6093 -0.1394 -0.0003 0.6153 -0.0078 -0.0002 XFOIL Viscous 0.5322 -0.1225 0.0055 0.5637 0.0027 0.0081

Table 1 Comparison Of Panel Code Results Table 1 compares the results for the target, converged, XFOIL inviscid, and XFOIL viscous solution. It can be noticed that the XFOIL inviscid solution falls almost exactly on the converged in viscid solution. Also, both inviscid solution are very close to the target indicating that the final geometry converged well. The viscous solution is reasonably close, but, as seen from table, the lift coefficient drops by several hundredths. This is because XFOIL accounts for the displacement thickness of the boundary layer when computing the viscous solution, which alters the panel code results accordingly.

4. Wind Tunnel Testing of Airfoils More Than Just Wall Corrections, By Peter Fuglsang, Stefane Bove.
Summary: This paper reports an experimental investigation on how to obtain 2D airfoil data in the LM Glasfiber wind tunnel. An experiment was designed with two airfoil models of different chord having the same shape. The experiment was used to investigate the influence of model size on wall corrections. Furthermore the influence from 3D flow for a typical 2D airfoil test setup was assessed together with the unambiguity of results from different measurement

methods. The applied wall corrections were verified and results were homogenous for Reynolds numbers of 3 and 6 million obtained from different combinations of chord and flow speed. Comparisons of pressure tap distributions at different span locations showed two dimensional flows for attached flow, whereas the flow was clearly three dimensional when the flow was separated close to maximum lift and in post stall. Comparison of lift obtained from airfoil pressure and wall pressures respectively showed good agreement for attached flow, but some differences for separated flow. Results for drag showed dependency on time and span averaging. The results showed that two dimensional results were feasible but that knowledge of the measurement method is essential when evaluating wind tunnel measurements. A benchmark against measurements in other wind tunnels showed that the small remaining uncertainties on wall corrections were insignificant compared to the variation of results between wind tunnels. Introduction: Wind tunnel testing of airfoils for wind turbines has become increasingly popular within the wind tunnel industry. An increased need for experimental verification has appeared as a follow of the trend to tailor for operation on specific wind turbine blades as opposed to previously where airfoils were predominantly selected among a limited variety of existing families. The continuous optimization of new blades occurs to increase efficiency as well as to move the barrier for how long blades can be for both existing and new turbine platforms. The tailoring of airfoils opens up further design optimization but also moves us away from our knowledge base. Numerical predictions have not yet reached a level where the accuracy is sufficient to eliminate the need for experimental verification. Wind tunnel testing is also motivated from the need to have accurate blade design data that are obtained at test conditions matching the operation conditions of the blade on the turbine. Due to the large size of today`s wind tunnel blades the number of suitable wind tunnels is limited with only little availability of free time slots. This fact together with high ambitions to improve the aerodynamics of wind turbine blades through extensive wind tunnel testing motivated LM Glafiber to built a new wind tunnel (LSWT) from scratch meeting up to date requirements to testing conditions and testing sections flow quality. It is tempting to use a model with a large chord and a high flow speed to match the Reynolds number of today s large blades. However, wall corrections due to blockage and interference become inaccurate. There are also other uncertainties from the three dimensional flow, which appears when the flow on the upper surface starts to separate causing uncertainty on the lift in post stall. Also the possible 3D variation of the flow in the span direction of the airfoil model causes uncertainty on drag. This paper reports an experimental investigation on how to obtain 2D airfoil data which was done as a part of the run-in of the LSWT facility. A special experiment was designed with two airfoil models of the NACA 64-618 shape having different chords of 600 mm and 900 mm. The model span of 1.35 m spanned the test section which has a height of 2.70 m. The airfoil

forces were derived from different measurement principles typically being used in similar 2D test setups. Two values were obtained for the airfoil lift coefficient from floor and ceiling wall pressures and airfoil pressure tabs respectively. The airfoil drag coefficient was obtained from wake rake measurements. Conclusion: Two airfoil models of the NACA 64-618 shape having different chords of 600 mm and 900 mm were tested in the LM Glasfiber LSWT wind tunnel at flow conditions corresponding to Reynolds numbers at 3x106 and 6x106. The results verified the applied classical wall corrections but also revealed that 2D airfoil testing is not trivial and more than just applying wall corrections. For appropriate interpretation of measured results in post stall it is important to know how the measurements were carried out and how the model aspect ratio affects the results. The investigation verified the test setup of LSWT where high quality end results for 2D airfoil data could be obtained and that results until Reynolds numbers of 6x106 are fully feasible. An investigation of the span variation of the 600 mm chord airfoil model pressure showed that the flow was two dimensional for angles of attack with attached flow, whereas measurements around maximum lift and in post stall showed three dimensional flow with a single stall cell in the centre region of the airfoil model. Results from the 900 mm chord model did not show the stall cell making the presence of this depended on the model aspect ratio. The span variation of drag was measured to be significant making it necessary to average drag in both time and space and furthermore to make sure that areas of the model with surface imperfections or pressure taps should not be present in the measurements. Results for lift obtained from airfoil and wall pressure respectively were in very good agreement for attached flow but were slightly different in the post stall making it necessary to know the measurement method when evaluating wind tunnel results. By comparing results from different model chords at the same Reynolds number, 3x106, it was possible to verify the applied wall corrections. The lift curve slopes as well as maximum lift and minimum cd values were consistent. However, the verification could only partially be extrapolated to Re = 6 x106, where a small difference in lift curve slope was found. A Benchmark of LSWT results against other wind tunnels put the verification of wall corrections done in this paper into perspective. The small uncertainties that remained from this study should be looked at together with the variation in results from different wind tunnels. Clearly, the uncertainty from comparing results between different wind tunnels is more significant that the remaining uncertainty from wall corrections.

5. Analysis, Design and Testing of Airfoils for use at Ultra-Low Reynolds Numbers By Peter J. kunz and Ilan M. Kroo.
The flight regime of micro-aircraft poses difficulties for aerodynamic analysis and design, but little experimental or computation work exists for aerodynamic surfaces operating at very low Reynolds numbers. Technological advances in micro-fabrication techniques and in the miniaturization of electronics are beginning to make mechanical micro-flight vehicles feasible from a system and manufacturing standpoint. Very small devices can now be built and there is numerous potential applications for these vehicles, but first they must be capable of flight. The reduced scale and low flight speeds of these vehicles result in Reynolds numbers on the order of 103. Although insects have been flying happily under these conditions for quite some time, this is a new flight environment for man-made aircraft. Aerodynamics at these Reynolds numbers is considerably different from those of more conventional aircraft. The flow is laminar and viscously dominated. Boundary layers are quite thick, often reaching a significant fraction of the chord length. Flow separation is an issue, even at low angles of attack. There is considerable literature on biological flight mechanisms, but there is very little detailed aerodynamic research available. The study of flight under these conditions is only now becoming more than an academic problem. A lack of suitable manufacturing technologies, the absence of sophisticated computational analysis methods, and the difficulties associated with accurate experimental work at this scale have restricted research in this area. Advances in technology have reduced the significance of the first two issues, but experimental work remains problematic. A reasonable solution is proposed in this paper. A study of two-dimensional airfoil geometry is a reasonable starting point for exploration. The parameters investigated include thickness, camber, and leading and trailing edge shapes. The airfoil sections are analyzed using a two-dimensional, incompressible, NavierStokes solver. The flow field is assumed to be fully laminar. Compared to performance at higher Reynolds numbers, the study demonstrates an order of magnitude increase in the drag and a similar sized reduction in lift to drag ratios. Although the drag rapidly increases with reductions in Re, significant lift coefficients are still attainable. Within the Re 103 range, reducing the Reynolds number results in an increase in the maximum steady-state lift. As Reynolds number is lowered, there is an alleviation of the leading edge suction peak which results in less adverse gradients along the low-pressure side of the airfoil. This delays separation and allows operation at higher angles of attack. Examination of the geometry parameters reveals several trends. There is an expected drag penalty associated with increased thickness, but also a significant reduction in the lift curve slope. An additional increment in the pressure recovery, due to thickness, degrades the lifting performance and results in an earlier onset of separation. Given the benefits of reduced airfoil thickness, the remaining studies utilize 2% thick geometries. The performance of these thinner airfoils appears to be insensitive to the thickness distribution, but the magnitude and distribution of camber are still highly effective parameters. Building on the results of the

geometry survey, several sections have been developed using an automated optimization method. Maximizing the two-dimensional lift to drag ratio at a given Reynolds number is the design goal. Validation of the computational results will consist of experimental flight testing of small gliders. These tiny aircraft are constructed with a specific airfoil of interest and designed to achieve a steady-state glide at a defined Reynolds number. This is full-scale flight testing at a very small size. Estimates of the aircraft s performance have been developed using the results of the two-dimensional section analyses. Comparison of the predicted performance with flight test data will provide a good assessment of the accuracy of the two-dimensional computational results.

6. High Lift low Reynolds Number Airfoil Design By Michael S. Selig and James J. Guglielmo , University Of Illinois at Urbana.
Introduction: A new high-lift airfoil design philosophy has been developed and experimentally validated through wind-tunnel tests. A key element of the high-lift design philosophy was to make use of a concave pressure recovery with aft loading. Three codes for airfoil design and analysis (PROFOIL, the Eppler code, and ISES) were used to design the example S1223 high-lift airfoil for a Reynolds number of 2 x 105. In wind tunnel tests, the new airfoil yielded a maximum lift coefficient of 2.2. With vortex generators and a 1% chord Gurney ap (used separately), the CLmax increased to 2.3. The airfoil demonstrates the rather dramatic gains in CLmax over those airfoils previously used for high-lift low Reynolds number applications. Increased payloads, shortened takeoff and landing distances, reduced aircraft noise, and lowered stall speeds can all be derived from the beneficial effects of improved high-lift airfoil aerodynamics. It is, therefore, not surprising that the classic problem of high-lift airfoil design has been and remains a topic of considerable interest. The purpose of this paper is to present a high-lift airfoil design philosophy for the increasingly important low Reynolds number regime in which small unmanned aerial vehicles (UAVs) operates. Only single element airfoils are considered in the current work. Airfoils for such aircraft typically operate in the Reynolds number range 2 X 105 to 5 X105. For example, U.S. Navy electronic warfare UAVs (e.g., LAURA5 and FLYRT aircraft) fl y at ship-like speeds ranging from 25 to 40 kn with payload requirements varying from 10 to 25 lb. The small vehicle size required for efficient shipboard storage coupled with low flight speeds and demanding payload requirements places great emphasis on high-lift low Reynolds number aerodynamics. A similar- sized aircraft, the hand-launched Pointer UAV operated by the U.S. Army is used to perform short-range reconnaissance missions. Moreover, small payload-laden UAVs have been envisioned for missions that involve atmospheric sampling, border surveillance, forest fi re detection/tracking, ship- or aircraft wreck survivor search and weather monitoring. In each case, high-lift airfoil performance can to varying degrees play an important role. Although not all of these airfoils were specially designed for high-lift, a predictable and anticipated trend emerges, the lower the Reynolds number, the lower the maximum lift. In particular, in going from a Reynolds number of 1X106 to 1X105, a sharp drop in

Cl,max is seen in the available data. The lower end of this range is of interest in the design of small UAVs based on current trends. In particular, this paper focuses on high-lift airfoil design for a Reynolds number of 2X105. High lift is rarely the only desirable feature of an airfoil. The airfoil lift-to-drag ratio, endurance parameter, thickness, pitching moment, stall characteristics, and sensitivity to roughness are all important factors, among others, that must each be weighed separately when one considers selecting or designing an airfoil. This study focuses on those factors most related to enhanced high-lift low Reynolds numbers airfoil performance. Conclusion: As a result of this work, it is clear that low Reynolds number airfoils can be designed to achieve lift coefficients much higher than previously thought possible. Such high-lift performance can be achieved through the use of a design philosophy that fully exploits the favorable effects of both a concave pressure recovery and aft loading. Application of this philosophy was demonstrated through the successful design of an airfoil that achieved a CLmax = 2.2 at a Re = 2X105. Surprisingly, as the example airfoil illustrates, the pressure recovery for this class of airfoils, though concave and close to a Stratford distribution, can be tailored to produce acceptable stall characteristics for UAV applications.

7. Optimization Of 2-D Flap Geometry Using Matlab and FUN3D By Gregory D. Howe
Introduction: This paper describes work done in the process of creating a workable system for the optimization of two-element high-lift airfoil design based on a fixed "cruise configuration" baseline. Methods were developed to define airfoil flap geometry, automatically create and run unstructured computational meshes based on this geometry, and to iteratively optimize this geometry. Validation cases are presented based on different optimization algorithms and parameters. Significant work is also presented on the attempt to characterize the design space of this problem in order to better understand the performance of different optimization routines. One of the more difficult problems in aerodynamic design is that of high-lift aerodynamics. This is largely because computational fluid dynamics has only recently reached the point where reliable CLmax computations can be done in a reasonably short amount of time. Accurate modeling of viscous flow phenomena is necessary not only to predict stall, but also to predict aerodynamic coefficients at angles-of-attack immediately preceding stall. Before the 1990s, the only CFD codes available to quickly compute lift coefficients near stall depended largely upon empirical methods to work the viscous part of the solution. Numerically optimizing these sorts of codes is dangerous, as the optimizer tends to get stuck in the assumptions of the empirical methods. There are several intermediate objectives that must be completed in order to have a reliable scheme for the optimization of flap CLMAX. The major components include:

y y y

Specification of airfoil and flap geometry , Automation of airfoil grid generation , Choice (and possibly design) of one or more optimization algorithms.

Conclusion: While the short conclusion is that the optimization process does in fact improve CLmax, there is still much work to be done. Certainly the simple robustness tests need to be performed by attempting to use the methodology on different base airfoils, at different flap deflections, and so on. There are a few more particular questions that need to be answered, though. The result that all of the cases run at 8 angle-of-attack experience some separation has called into question the decision to run the optimizers there. A course of action is unclear, however. Should the optimizer be run at 6, where the airfoil does not seem to be experiencing separation? Or should the highest CLmax be sought by running them at 11, closer to CLmax of the original airfoil? There is also the possibility that the optimizer could be designed to try to avoid these separated cases, possibly attempting to maximize CL/ CD or something more unusual like CL2/ CD. The flap leading edge geometry definitions remain somewhat unproven. It is clear that the methodology can create a flap that subjectively looks like a flap should, but whether the method is general enough to describe the best possible flap for an airfoil is unknown. Future tests should include cases for which known flaps have been developed and tested, so that there can be a comparison between optimizer results and the traditional methods of developing a flap geometry. Additionally, all of the optimization cases run so far have used 4th-order Bezier curves to define the flaps. This was chosen simply because it is the simplest case. It is certainly possible that the greater degree of freedom created by adding more control points would allow for better solutions.

8. Flight- Determined Subsonic Lift and Drag Characteristics of Seven Lifting-Body and Wing-Body Reentry Vehicles Configurations with Truncated Bases, By Edwin J. Saltzman, K. Charles Wang, Kenneth W. Iliff.

Introduction: This paper examines flight-measured subsonic lift and drag characteristics of seven lifting-body and wing body reentry vehicle configurations with truncated bases. The seven vehicles are the full-scale M2-F1, M2-F2, HL-10, X-24A, X-24B, and X-15 vehicles and the Space Shuttle prototype. In recent years, interest has been renewed in controlled reentry from low-Earth orbit and the Earth s upper atmosphere. This interest has been motivated by several factors: a growing commercial space launch market and its desire for a low-cost, reusable means of space access; the need for a crew return/rescue vehicle from the International Space Station; and the potential for future military space operations. Because of the current interest in lifting reentry shapes, this paper reexamines lift and drag`s characteristics of the seven aforementioned vehicles during subsonic unpowered flight, and presents a unified analysis of their subsonic aerodynamic performance that enables meaningful comparisons with new lifting reentry

designs. The vehicles examined in this paper, the M2-F1, M2-F2, HL-10, X-24A, X-24B, and X-15 vehicles and the Shuttle prototype, comprise a unique class of aircraft. Summary: Despite the success of the X-15 unpowered landing experience, the early planning for the Space Shuttle included pop out auxiliary engines to ensure safe horizontal landings. Thompson, an X-15 and lifting body research pilot, argued that the X-15 and lifting body experience rendered landing engines for the Space Shuttle as an unnecessary weight and payload penalty. Currently, new lifting reentry vehicles are being developed for rescue missions from space and to serve as reusable launch vehicles. These vehicles have much in common with the lifting bodies described herein and, if aspect ratio is increased somewhat, with the X-15 aircraft and the Shuttle prototype. This section assembles methods and metrics (performance parameters) used in the analysis of the subject lift and drag data. The primary metrics of aerodynamic performance include lift-curve slope; a modified Oswald lifting-efficiency factor; the drag-due to- lift factor; maximum lift-to-drag ratio; and for minimum drag analysis, equivalent parasite drag area, equivalent skin friction coefficient, base pressure coefficient, base drag coefficient, and fore body drag coefficient. The most exact theoretical solution for unswept, rectangular wings at incompressible conditions is considered to be that derived by Krienes. Krienes relationship for lift-curve slope and aspect ratio, A, is well-represented by the following relationship from Helmbold as expressed by Polhamus:

At the lowest aspect ratios (A<1) , equation (1) merges with the linear relationship of Jones,27 which follows:

Equations (1) and (2) represent lift due to circulation. The metrics used to evaluate the lift-related drag of the subject vehicles are the drag-due-to-lift factor; and the modified Oswald lifting efficiency factor, which is a measure of the span wise distribution of lift. The Oswald factor as applied herein has been modified as proposed by Wendt:

In this modified form of Oswald s efficiency factor, CLmin and CDmin are the values of lift and drag coefficient at the vertex of the parabolic or nearly parabolic relationship of CL as a function of CD (that is, the drag polar), which does not necessarily occur at zero lift. This

condition exists for five of the seven vehicles considered in this study. Both lift-related drag factors represent lift coefficients extending to greater than that required to obtain maximum lift-to-drag ratio. The value (L/D)max achieved by each of the subject vehicles at subsonic speeds is presented as a function of b2/AW. This form of aspect ratio is referred to as the wetted aspect ratio. This presentation includes a reference framework consisting of a family of curves representing constant values of equivalent skin friction coefficient, which is a form of minimum drag coefficient, (which includes both fore body and base drag). Thus, if

Then

Although CFe is called the equivalent skin friction coefficient, the operative word is equivalent because CFe contains base drag, separation losses, protuberance drag, and other losses in addition to skin friction. The family of reference curves is analogous to that employed by Stinton, and the curves are defined by the following often-used expression from Loftin:

Minimum drag is considered in several formats. When the lift coefficient and drag coefficient are based on vehicle plan form reference area, the minimum drag coefficient can be defined as noted earlier in equation (4). The discussion on maximum lift-to-drag ratio also revealed that another metric for minimum drag coefficient is the equivalent skin friction coefficient (eq. (5)), which is obtained by basing the minimum drag coefficient on the wetted area, AW . The wetted area for each vehicle is considered to be the wetted area of the respective fore body, which includes the body and wings or fins, and is thus the sum of all outer mold-line surfaces ahead of an associated base or trailing edge. Another format for comparing minimum drag for various configurations is called the equivalent parasite drag area, f.

Mathematically speaking, the following exists:

Where CDfore,s is the fore body drag coefficient is referenced to S,CFe is the equivalent skin friction coefficient due to fore body only, CDb is the coefficient of base drag, and Ab is the base area. Significant fore body drag losses exist in addition to the losses caused by skin friction alone. A way to quantify the sum of these losses is to compare the measured minimum drag of a vehicle with the sum of the measured base drag and the calculated skin friction drag for completely attached, turbulent, boundary-layer flow. The difference that results from this comparison represents losses from multiple sources, which are designated excess fore body drag. The calculated, idealized, sum of the base drag and skin friction drag for each vehicle is obtained from:

Where CF is the turbulent skin friction coefficient (calculated) of the fore body and c is a base pressure profile factor. A common practice by wind-tunnel and flight experimenters has been to define a base drag coefficient increment as:

where is obtained from a few scattered pressure measurements within the confines of the base surface. Thus, equation (11) is based on the assumption that the base pressure profile (consisting of the average of the pressures measured within a specific base region) was flat to the very edge of the base. However, the pressure profile is known to be somewhat rounded along the edge. Flight-measured base pressure coefficients, base pressure coefficients derived from published incremental drag attributed to the base, and estimated base pressure coefficients derived from those of a closely related, after body-base configuration are compared with two analytical equations developed by Hoerner. These equations were derived from wind tunnel experiments of small-scale models. Hoerner s equation for three-dimensional axisymmetric bodies of revolution is as follows (where K=0.029 ):

Hoerner s equation for quasi-two-dimensional base flow conditions that generate the wellknown Karman vortex street is:

The flight-measured lift and drag coefficients for all seven vehicles were obtained by the accelerometer method. The operative relationships for subsonic unpowered gliding flight are:

Where an and al are the normal and longitudinal accelerations in g units, is the angle of attack, W is the vehicle weight, q is the free-stream dynamic pressure, and S is the reference area. Results and Discussion: The results of the current study are presented and discussed under four subheadings: Lift-Curve Slope, Lift-Related Drag, Lift-to-Drag Ratio, and several metrics of Minimum Drag. Table 2 shows the data to be considered as derived from their respective references. The CL and aspect-ratio values shown are, of course, subject to the values of the reference area, S, that were used in the various referenced documents. Use of the proper reference area and span is important towards achieving some understanding of how the lifting characteristics of the various configurations relate to each other, to generic wind-tunnel model data, and to theory. Figure 2(a) shows the lift-curve slope data (the solid symbols) for five of the seven vehicles as published in the respective references plotted as functions of aspect ratio. Figure 2(a) also shows the relationships of CL to aspect ratio as defined by Helmbold (eq. (1)) and, for the lowest aspect ratios, the linear relationship of Jones (eq. (2)). Neither of these relationships accounts for lift from vortices generated by sharp, highly swept leading edges.

The lift-curve slopes for each of the flight vehicles were expected to occur below the Jones and Helmbold relationships, which represent maximum efficiency for medium- or low-aspect-ratio configurations that obtain their lift from circulation. However, the results from both M2 vehicles and the X-24A vehicle, as shown by the solid symbols in figure 2(a), considerably exceed these expectations. In addition, the X-15 solid-symbol lift-curve slope greatly exceeds the Helmbold relationship. Representative reference areas have been assigned for five of the seven vehicles; the other two vehicles were already assigned representative reference areas, as published. The revised reference areas and the resulting lift-curve slopes are also shown in table 2. Figure 2(a) also shows low-aspect-ratio wind-tunnel model results.

At high- and moderate-aspect ratios, lift-curve slope is diminished by wing sweep. At aspect ratios less than 2, however, the influence of sweep on lift is weak. Figure 3, shows this characteristic. Consequently, wing-sweep effects for the X-15 aircraft and the Shuttle prototype have not been addressed in this discussion of lift-curve slope.

as drag-due-to-lift factor plotted as a function of the reciprocal of aspect ratio (1/A). Included as a reference framework is a family of lines representing the theoretical relationship for an ideal elliptical span loading, wherein =1 , and for significantly less optimum load distributions represented by <1, which are expected for the vehicles reported here. For several of the subject vehicles, the minimum drag did not occur at zero lift. For these vehicles, their polars were displaced, CDmin and occurred at some finite lift coefficient defined as CLmin, which is the lift coefficient at the vertex of the parabolic, or nearly parabolic, polar. Note that only the winged vehicles, the X-15 aircraft and Shuttle prototype, have drag-due-tolift factors below this line, although one configuration of the X-24A vehicle is borderline. Figure 4 shows a qualitative interpretation of the relative lifting efficiency of the subject vehicles. All slopes of shown in figure 4 and in table 3 are based on the revised reference areas used and discussed in the Lift-Curve Slope section (table 2). The lifting-efficiency factor is not influenced by the choice of reference area.

Summary of Results: Flight-determined lift and drag characteristics from seven blunt-based lifting-body and wingbody reentry configurations have been compared and related to several standards of aerodynamic efficiency. For liftcurve slope, limited comparisons are made with generic windtunnel model results and the theoretical relationships of Jones and Helmbold. A summary of major results is as follows: I. Base pressure coefficient data from the X-15, the M2-F3, and the Space Shuttle vehicles indicate that Hoerner s equation relating base pressure to threedimensional forebody drag requires a larger numerator coefficient in order to represent largescale flight vehicles. A tentative range of values for the numerator coefficient is from 0.09 to 0.10 rather than 0.029, which is based on small-scale model data. Evidence exists that subsonic flow separating from a relatively large, sharp-edged three dimensional base can exhibit quasi-two-dimensional characteristics and base pressure coefficients. The nature of the Hoerner base-pressure-to fore body-drag relationship (regardless of whether his three-dimensional or two-dimensional equation is used, or the numerator coefficient value) causes base drag and fore body drag to combine to form an optimum minimum drag (a drag bucket ) over a small range of fore body drag. The magnitude of fore body drag coefficient that defines the bucket depends on the ratio of base area to wetted area of the respective vehicle. Conversely, a strong relationship between forebody drag and fineness ratio (favoring, of course, the higher fineness ratios) has been demonstrated to exist. This characteristic, in concert with the possibility of achieving the aforementioned drag bucket, underlines the importance of obtaining more large-scale free-flight base pressure and forebody drag data. Minimum equivalent parasite drag area values for the vehicles range from 6.5 ft2 to 164 ft2. Division of equivalent drag area by the associated wetted area provided equivalent skin friction coefficients ranging from approximately 0.009 to 0.020, excluding the less efficient body-flap configurations (these coefficients include base drag). The M2-F2 data demonstrate that the lift-curve slope of very-low-aspect-ratio lifting bodies can exceed the lift-curve slope values represented by the relationships of Jones or Helmbold for aspect ratios less than approximately 1. Excepting the M2-F1 and the HL-10 vehicles, the remaining five vehicles form an array (a band of over a range of ) that should be a useful reference source against which to relate future reentry-type vehicles.

II.

III.

IV.

V.

VI.

VII.

9. Computer Program To Obtain Ordinates for NACA airfoils, By Charles L.Ladson, Cuyler W. Brook Jr, and Acquilla S. Hill, Langley Research Center, Hampton , Virginia. December 1996.
Introduction: Computer programs to produce the ordinates for airfoils of any thickness, thickness distribution, or camber in the NACA airfoil series were developed in the early 1970 s and are published as NASA TM X-3069 and TM X-3284. For analytic airfoils, the ordinates are exact. For the 6-series and all but the leading edge of the 6A-series airfoils, agreement between the ordinates obtained from the program and previously published ordinates is generally within 5 10-5 chord. Although modern high-speed aircraft generally make use of advanced NASA supercritical airfoil sections, there is still a demand for information on the NACA series of airfoil sections, which were developed over 50 years ago. Computer programs were developed in the early 1970 s to produce the ordinates for airfoils of any thickness, thickness distribution, or camber in the NACA airfoil series. These programs are published in references 1 and 2. These programs, however, were written in the Langley Research Center version of FORTRAN IV and are not easily portable to other computers. The purpose of this paper is to describe an updated version of these programs. The goal was to combine both programs into a single program that could be executed on a wide variety of personal computers and workstations as well as mainframes. This program is written in ANSI FORTRAN 77 and can be compiled to run on DOS, UNIX, and VMS based personal computers and workstations as well as mainframes. Summary: The traditional NACA airfoil designations are shorthand codes representing the essential elements (such as thickness-chord ratio, camber, design lift coefficient) controlling the shape of a profile generated within a given airfoil type. NACA 4-digit-series airfoils. Symmetric airfoils in the 4-digit-series family are designated by a 4-digit number of the form NACA 00xx. The first two digits indicate a symmetric airfoil; the second two, the thickness-chord ratio. Ordinates for the NACA 4-digit airfoil family are described by an equation of the form:

To define an airfoil in this family, the only input necessary to the computer program is the desired thickness chord ratio. NACA 4-digit-modified-series airfoils. The 4-digitmodified- series airfoils are designated by a 4-digit number followed by a dash and a 2-digit number (such as NACA 0012-63). The first two digits are zero for a symmetrical airfoil and the second two digits indicate the thicknesschord ratio. The first digit after the dash is a leading-edge-radius index number, and the second is the location of maximum thickness in tenths of chord aft of the leading edge. The design equation for the 4-digit-series airfoil family was modified so that the same basic shape was retained but variations in leading-edge radius and chord wise location of maximum thickness could be made. Ordinates for these airfoils are determined from the following equations:

from leading edge to maximum thickness, and

from maximum thickness to trailing edge. NACA 16-series airfoils. The NACA 16-series airfoil family is described in references 6 and 7. From the equation for the ordinates in reference 7, this series is a special case of the 4digit-modified family although this is not directly stated in the references. The 16-series airfoils are thus defined as having a leading-edge index of 4 and a location of maximum thickness at 0.50 chord. The designation NACA 16-012 airfoil is equivalent to an NACA 0012-45. Program Capabilities: AIRFOLS is the computer program which was developed to provide the airfoil shapes. AIRFOLS, the NACA Airfoil Ordinate Generator, is the result of merging two previous efforts reported. This program calculates and then reports the ordinates and surface slope for airfoils of any thickness, thickness distribution, or camber in the designated NACA series. AIRFOLS is a portable, ANSI FORTRAN 77 code with limited platform dependencies. Parameters describing the desired airfoil are entered into the software with menu and prompt driven input. Output consists of a report file and a data file. Provisions have been made in the AIRFOLS program to combine basic airfoil shapes and camber lines from different series so that nonstandard as well as standard airfoils can be generated. The analytical design equations for both symmetrical and cambered airfoils in the NACA 4-digit-series, 4-digit-series modified, 5-digitseries, 5-digit-series modified, 16-series, 6-series, and 6A-series airfoil families have been implemented. The camberline designations available are the 2-digit, 3-digit, 3-digit-reflex, 6-series, and 6A-series. The program achieves portability by limiting machine-specific code. Conclusion: All the airfoils and camber lines generated by this program are defined by closed analytical expressions and no approximations have been made in the program. Thus, all results are exact. Many cases have been run and compared with previously published results to check the procedure, and for all cases the comparisons were exact except for occasional differences beyond the fifth digit caused by rounding differences. Computer programs to produce the ordinates for airfoils of any thickness, thickness distribution, or camber in the NACA airfoil series were developed in the early 1970 s and are published as NASA TM X-3069 and TM X-3284. For analytic airfoils, the ordinates are exact. For the 6-series and all but the leading edge of the 6Aseries airfoils, agreement between the ordinates obtained from the program and previously published ordinates is generally within 5 10-5 chord. Since the publication of these programs, the use of personal computers and individual workstations has proliferated. This report describes a computer program that combines the capabilities of the previously published versions. This program is written in ANSI FORTRAN 77 and can be compiled to run on DOS, UNIX, and VMS based personal computers and workstations as well as mainframes. An effort was made to make all inputs to the program as simple as possible to use and to lead the user through the process by means of a menu.

10. Validation of a Navier-stokes Solver For Airfoil High-Lift Analysis, By Marc Langlosis and Farzad Mokhtarian , Advanced aerodynamics Dept; Bombardier Aerospace, Montereal Canada.
Introduction: This paper presents the results of validation exercise of a two-dimensional NavierStokes solver conducted for high-lift application by Bombardier Aerospace. The aim of the study was to investigate the capability of the code to accurately predict the complex flows around high-lift configurations and the requirements in terms of mesh density and turbulence modeling for successful computations. The experimental data for this validation comes from a dedicated two-dimensional high-lift test conducted in January 2001 in the Canadian Institute for Aerospace Research (IAR) high-Reynolds number 15 in X 60 in. test facility. Three test cases have been selected from the experimental data: a clean airfoil, a slat-deployed configuration and a flapdeployed configuration. The parameters that were investigated included the extent of the computational domain, the wall spacing, local mesh refinement and modeling. As the complexity of the high-lift systems increased, increasingly sophisticated CFD analyses were used to design efficient multi-element airfoils and wings. For two-dimensional and three-dimensional analyses, current methods cover the whole range from panel method to Navier-Stokes

solvers.
While viscous panel methods and Euler/boundary-layer codes are still the main tool for wing design, Navier-Stokes codes are routinely used today for airfoil design. However, the requirements in terms of discretization as well as transition specification and turbulence modeling for accurate predictions of maximum lift (CL max) and lift-to drag ratio (L/D) still remain to be clearly established. Summary: The testing was conducted in the Institute for Aerospace Research (IAR) high-Reynolds number 5'x5' trisonic wind tunnel located in Ottawa. This blow-down tunnel has a 15 x60 twodimensional test section. It can operate at unit Reynolds numbers up to 40x106 per foot. NSU2D is an unstructured-grid compressible 2-D solver developed by D.J. Mavriplis. It solves the Reynolds-averaged Navier-Stokes equations using a Galerkin finite element approach. Flow and turbulence variables are stored at the vertices of the mesh. The convective fluxes are computed at the vertices and assumed to vary linearly of the triangular elements, while the velocity gradients in the viscous stresses are computed at the centers of the triangles with the flow variables assumed to vary linearly over the triangles. The discretized mean flow equations are integrated in time using an explicit five-stage Runge-Kutta scheme devised to ensure rapid damping of high-frequency errors. Convergence acceleration techniques include local time stepping and implicit residual smoothing as well as the use of an algebraic multi grid algorithm. The influence of some basic mesh and solver parameters was first investigated on a clean airfoil configuration at a relatively high incidence of 14 degrees. Clearly, we need to establish the capability of the solver to predict stall and maximum lift. However, stall is triggered by flow separation, which, even on a nominally two-dimensional test configuration is a threedimensional phenomenon. Inspection of the span wise pressure distributions near the airfoil trailing edge shows that the 2-D nature of the flow is often lost before maximum lift is attained,

and the margin increases with slat and flap deflection. Figure 2 shows an example of span wise pressure distributions and the loss of two-dimensionality with incidence.

The dependence of the solution on computational domain size is illustrated in Figure 3, which shows the variation of the lift coefficient with the distance of the far-field boundary from the airfoil, expressed in airfoil chords. Even though NSU2D incorporates a circulation correction on the far field, it is clear that a 10-chord distance for the far-field boundary is not enough to guarantee a numerically accurate solution. Figure 4 shows the pressure distributions obtained with this spacing and spacings 10 times smaller and larger. Whereas decreasing the spacing does not change the solution, thereby showing that the 10-6 spacing is indeed adequate, increasing it results in a lower suction peak and a small trailing edge separation bubble which is otherwise not present. The normal spacing has an even stronger influence on the skin friction distribution, as shown in Figure 5, and hence on the drag prediction. The results presented so far were obtained using Menter s baseline k-w turbulence model. Figure 6 shows the influence of the turbulence model on the pressure distribution for this test case. The pressure peak on the airfoil seems to be slightly better predicted with the one-equation Spalart-Allmaras turbulence model, but this difference is very small. On the other hand, the k-w model matches the experimental trailing edge pressure distribution better than the one-equation model, which shows a separation bubble. It should be noted that this bubble does not disappear when the SpalartAllmaras model is used on the mesh with the finer wall spacing.

Conclusion: Selected comparisons of two-dimensional Navier-Stokes predictions with the data from a 2-D high-lift wind tunnel test have been presented. The aim of this study was to investigate the ability of the NSU2D Navier- Stokes solver to properly predict the complex flow phenomena around multi-element airfoils and to establish some guidelines for its successful application. Three test cases were used in this study, consisting of a clean airfoil, a slat-deployed configuration and a flap-deployed configuration. Regarding mesh density, the following recommendations can be made. For the same surface density, the computational domain should extend at least 20 to 30 chords from the airfoil surface to insure a mesh-converged solution. For Reynolds numbers of the same order as that considered here, mesh spacing in the normal direction of 10-6 times the chord is sufficient to obtain an adequate value of y+ of the order of 1. Using larger wall spacing results in the erroneous prediction of a trailing separation bubble on the clean airfoil and much reduced skin friction coefficients near the leading edge. Mesh refinement based on curvature alone may not be sufficient to capture all flow features. In particular, user-specified local mesh refinement is necessary on the surfaces in the recirculation regions. Volume refinement through interior point sources does not seem however to provide a significant additional improvement in this case. The modeling of a wake from the lower lip of the slat does not appear to improve the quality of the predictions. All of these points to the fact that mesh adaptation could be a very useful tool for accurate high-lift flow predictions. The selection of a quantity on which to adapt the mesh in separated flow regions may however pose a problem. Concerning turbulence modeling, the results presented point to the superiority of the two-equation k-w model over the one-equation Spalart-Allmaras model for high-lift applications. The latter tends to predict trailing edge separation ahead of the former and of the experimental measurements. With the slat deployed, the location of the main airfoil attachment point and the pressure on its forward lower surface and in the slat cove are also better predicted with the two-equation model.

11. CFD Analysis of Air Turbines as Power Take-Off Systems in Oscillating Water Column Wave Energy Conversion Plant, By A. Gareev, P. Cooper and P. B. Kosasih, University of Wollongong, AUSTRALIA.
Introduction: This paper presents the results of CFD simulations of reversing flow air turbines used as the power takeoff system in Oscillating Water Column (OWC) Wave Energy Conversion (WEC) plant. One of the simpler tools to analyze such turbines is the blade element/actuator disc methodology. This requires the input of interference factors to model how the lift and drag characteristics of the cascade of blades on the turbine rotor are related to those of a single isolated aerofoil. In the first part of the paper, CFD modeling to obtain the lift and drag characteristics of various airfoils arranged in linear cascades at different stagger angles is described. The CFD cascade lift and drag data are compared with reported experimental cascade aerodynamic data. The agreement within the range of usable stagger angles is excellent in the pre-stall range with some deviations shown in the post-stall. A comparison is also made

between our 2D CFD interference factors and those previously reported by Weinig and others who used analytical, in viscid flow theory. It is found that the Weinig in viscid flow theory provides a reasonable prediction of the lift interference factor providing that both the angle of attack is relatively low and that the thickness of the blades is relatively small compared to the distance between blades. In the second part of the paper, three dimensional simulations of a Wells air turbine rotor using CFD unstructured and structured grid designs are described. The results of the three-dimensional CFD simulations were then compared with those from our nondimensional blade element model incorporating the linear cascade aerodynamic data described in the first part of the paper. The two sets of results are compared in terms of torque coefficient and pressure coefficient. The overall capture efficiency of any OWC WEC is critically dependent on the performance of the air turbine used to extract mechanical energy from the air flow generated by the moving free surface in the OWC chamber (Fig. 1). This paper presents recent fundamental Computational Fluid Dynamics (CFD) research on the lift and drag characteristics of linear cascade aerofoil and aerodynamic behavior of flow through OWC air turbine rotors without guide vanes with either fixed or variable pitch blades.

Summary: One of the difficulties of the blade element model is that in a relatively high solidity rotor, as is common in many OWC air turbines, the interference between the blades in the circular cascade of blades is not known a priori. It is also true that fundamental experimental testing of a cascade of aerofoil at high stagger (or pitch) angles, , (see Fig. 2) is virtually impossible in a wind tunnel. In the past the interference of adjacent blades in a cascade at various stagger angles has been estimated by a number of methods including: i) potential flow analysis of flat plate aerofoil arranged in a straight-line cascade as proposed by Weinig and the Method of Singularities. It should be noted that both these in viscid theoretical methods make it possible to estimate cascade lift coefficients, but not the drag values. A third approach reported is a semi empirical method based on a correlation between computed values of mean aerodynamic force coefficients from a turbine test and 2D aerofoils data obtained in a wind tunnel .

Accurate prediction of turbine performance parameters using the blade element/actuator disc methodology requires the input of reliable lift and drag data for blades arranged in a cascade, as shown in Fig. 2. Relatively little fundamental research has been published on the flow in such rotors until recently. For the general case of a cascade with arbitrary solidity and stagger angle Weinig s in viscid flow analysis provides the following estimate of lift coefficient for an isolated aerofoil prior to stall modified by the interference factor, k0:

Where m is the angle of attack (based on the mean of the velocities upstream and downstream of the cascade) and where k0 (shown in Fig. 3) was found through the solution of a set of algebraic equations. In the case where the stagger angle = 0, which is applicable to the case of the Wells turbine, Weinig gives the interference factor as:

Equation (2) has been extensively used by researchers as the correction factor to modify the lift coefficient of fixed-pitch Wells turbine blades. It should be noted that Weinig s theory indicates that the interference factor is predicted to be a function only of solidity and independent of angle of attack, . However, there has previously been very little, if any, validation of this relationship for practical aerofoil cascades. The present authors have compared the in viscid flow results of (2) for a 2D linear and infinite cascade of NACA0012 aerofoil (with a stagger angle of 0) with CFD simulations using the ANSYS CFX code. For all simulations reported in this paper the CFX solver was run using the k- turbulence model with the high resolution advection scheme and auto timescale . The CFD results of interference factor as a function of the inverse of solidity, s/c, and the mean angle of incidence, m, are presented in Fig.4

The 2D simulations were carried out at Re = 7105 with an unstructured grid of ~2.0 x 106 elements. The average cpu time for each run was about 1 hour. It is seen from Fig. 4 that there was generally good agreement between Eqn. (2) and the CFD results for small angles of attack. However, the agreement is not good for m > 10 as a result of the in viscid flow theory not predicting the onset of stall, i.e. the CFD results show lower values of k0 than Weinig s results due to stall and the formation of a separation region on the suction side of the aerofoil which is not accounted for in Weinig s in viscid flow analysis (Fig. 5). However for high solidity cascades (low s/c) the interference effect between adjacent blades suppresses this separation region resulting in a steep rise in cascade lift coefficient and hence an increased interference factor. The blockage effect on the pressure side is demonstrated in Fig. 6. For blades of finite thickness, small values of s/c lead to an increase in velocity in the passages and an increase in lift. A further comparison, using a different cascade geometry, of the interference factor k0 predicted from Weinig s in viscid flow analysis (as shown in Fig. 3) with those of a fully viscous flow CFD analysis has been conducted on a cascade of blades with the same dimensions and pitch as the scale model of the Denniss-Auld axial flow, variable-pitch turbine described by Finnigan and Auld [5]. The blade profile used for this prototype turbine was based on the NACA 65-418 with maximum camber height of 6% and maximum thickness to chord ratio of 18%. The blade geometry was symmetric about the mid-chord and was formed by combining two front halves of the NACA 65-418. Values of lift interference factor k0 deduced from the CFD simulations for a constant upstream angle of attack = ( - ) = 10 as a function of stagger angle and solidity is shown in Fig. 7. The results for the drag interference factor 0 = CD/CD0 for the same cascade are also shown in Fig. 8. Two-dimensional CFD simulations have been carried out

by using an unstructured mesh having a total number of elements of between 1.2 to 1.45 million depending on the stagger angle. Cascade flow was modelled using the k- turbulence model. The average cpu time for each run was about 55 minutes. Note that the lift and drag coefficients for the cascades have been calculated using the mean of the upstream and downstream angles of incidence as defined by Weinig .

Conclusions: The CFD analysis has demonstrated that for angles of attack such that 10Weinig s in viscid flow analysis provides an accurate prediction of the interference factor for lift, k0. Weinig s analysis did not account for the finite thickness of practical blades in a cascade, and our CFD results indicate a higher interference factor for high values of rotor solidity because of higher air velocities within the blade passages than would occur for blades of zero thickness. The results of the CFD analysis of the effective drag interference factors, 0, for a cascade are also presented. It is demonstrated that single aerofoil data and interference factors may be used effectively in blade element/actuator disk analysis of a variable pitch Wells turbine. Our blade element results have been compared to both the original experimental results of Tease and with the full 3D CFD simulations of the present study. Comparison of the data shows that the blade element results and CFD match the non-dimensional pressure experimental data well, however the match in the case of the torque coefficient is less precise.

12. Dynamic stall Experiments on Oscillating Airfoils, By W. J. Mccroskey, L.W. Carr and K.W.McAlister, U.S. army Air mobility R&D Laboratory, Moffett Field, Calif.
Introduction: The phenomenon of dynamic stall on airfoils and lifting surfaces in unsteady-flow environments has been studied for many years, both as an important practical problem and as a challenging fundamental one as well. Within the past decade, it has been established that a predominant feature of dynamic stall is the shedding of a strong vortex-like disturbance from the leading-edge region, this vortex passes over the upper surface of the airfoil, distorting the chord wise pressure distribution and producing transient forces and moments that are fundamentally different from their static stall counterparts (as illustrated in Fig. 1). This unique viscous-in viscid interaction is not well understood at present, although it has been proposed that the process begins with the sudden bursting of a leading-edge laminar separation bubble. The present experimental investigation is an outgrowth of the oscillating airfoil study described by Martin et al. Although that previous study linked the measured transient pressure distribution to flow visualizations of the vortex shedding and passage over the airfoil, the measurements did not reveal the detailed nature of the boundary-layer separation. This paper describes the most important results that have been obtained from the first phase of the rather comprehensive experiment, including the different types of boundary- layer separation that were observed on different airfoils, the general effects of the vortex shedding phenomenon at different frequencies, and the effect of leading-edge geometry on the normal force and. pitching moment coefficients. Summary: The basic configuration was the same as that used by Martin et al a NACA 0012 airfoil of 1.22 m chord and 2.0 m span mounted vertically in the U.S. Army AMRDL 7x 10-ft wind tunnel. The model was oscillated sinusoid ally in pitch about the quarter chord. A number of leadingedge modifications (illustrated in Fig. 2) were attached to the basic NACA 0012 profile in an attempt to modify the dynamic-stall characteristics by altering the nature of the bubble. The ONERA "0012U Extension Cambre'" airfoil13 was chosen to study the effects of camber.

Results and Discussion: The lower part of Table 1 indicates the variations in the stall characteristics that were observed as the test parameters were varied one at a time. The general vortex-shedding phenomenon was observed for all the airfoil shapes tested in the present investigation, whether the boundary-layer trip was used or not. Thus, it appears that this distinguishing feature of dynamic stall is essentially independent of Reynolds number, airfoil geometry, and state of the boundary layer. However, major differences were observed in the stall angle, in the time history of the aerodynamic forces and moments, and in the detailed nature of the boundary-layer separation leading up to dynamic stall. As shown in the following sections, these differences were found to depend upon reduced frequency, upon whether the airfoil stalled due to the bursting of the leading-edge laminar separation bubble, and upon whether the airfoil stalled well before amax, or near the top of the cycle where a. 0.

The region of highly disturbed boundary-layer flow progresses upstream with increasing incidence, or time, to somewhere in the vicinity of X = 0.3. Small disturbances begin to appear in the surface pressure distribution slightly after incipient flow reversal is indicated by the hot wires; this is indicated by the open circles in Fig. 4. Generally, the tufts and smoke jets from the rear ports begin to reverse direction and move upstream somewhat later than when the first disturbances to the static pressure appear, and this defines the upper part of the flow visualization boundary shown in Fig. 4. At  t=1.0, or = 23.4, the boundary-layer flow on the front of the airfoil abruptly breaks down catastrophically, or separates. As near as can be determined from the hot-wire and static-pressure data, this event occurs simultaneously in the turbulent flow from X = 0.3Q almost to the leading edge, although the boundary flow in the immediate vicinity of the laminar leading-edge bubble does not completely separate until slightly later. When this abrupt separation occurs, the static suction at X= 0.10 begins to rise, indicating the initial formation of the vortex. As this vortex begins to form and move downstream, the magnitude of the reversed velocity near the surface of the model increases. The solid symbols in Fig. 4 indicate the negative peak in static pressure and a major peak in the reverse velocity as measured by the hot wires and rearward- facing pitot tubes; these indicate that the vortex is traveling at approximately 35-40% of free stream velocity.

Conclusion: Figures 13-15 summarize the results for the three different types of boundary-layer separation that were observed in this experimental investigation of dynamic stall on oscillating airfoils; y trailing-edge stall developing from a relatively gradual progression of boundary-layer flow reversal and separation, from the trailing edge toward the leading edge; y leading-edge stall caused by an abrupt breakdown of the turbulent flow on the forward portion of the airfoil, following an initial progression of flow reversal from the trailing edge, and y two forms of leading-edge stall due to the abrupt bursting of a leading-edge laminar separation bubble. The most important result of the experiment is that leading-edge bubble bursting is much more of a special case than had been commonly assumed heretofore. In the important case of dynamic stall on the NACA 0012 airfoil, the shed vortex appears to be fed its initial vortices by the abrupt, unsteady separation of the turbulent boundary layer, not the laminar bubble. Regardless of the type of boundary-layer separation, the main difference between static and dynamic stall is the vortex shedding phenomenon. Provided the flow separates over most or all of the upper surface of the airfoil, a vortex-like disturbance develops and passes over the airfoil at a convection velocity on the order of 1/3 to 1/2 of U. The transient aerodynamic forces and moments that are produced are considerably larger and different in nature from their steadystate counterparts. An exception was noted in cases where trailing edge stall did not penetrate all the way to the leading edge during any portion of the cycle, but even in these partial-stall cases, considerable hysteresis in CN and CM vs a. was observed. Although the vortex-shedding phenomenon appears to be a general feature of all oscillating airfoils over a wide range of Reynolds numbers, significant differences in the dynamic force and moment characteristics were observed on different airfoils. This suggests that future improvements in airfoil design would be possible if dynamic stall were better understood theoretically.

13. Aerodynamics Analysis of High-Lift Airfoils By A Dual-Time Integration Method With Upwind Splitting, By Yang-Yao Niu and Whey-Fone Tsait , National Center for High Performance Computing, Hsin-Chu, Taiwan, R.O.C.
Introduction: The high-lift system performance, a study of low speed aerodynamics, has a major influence on the design and safety of any airplane configuration. One of the high-lift systems is the high speed ground vehicle, such as a racing car, for which performance is enhanced by increasing the down force from multi-element wings mounted on the vehicle. The maximum negative lift can be produced by adding a short, 90 degree inverted flat plate at the trailing edge of the airfoils to increase aerodynamic down force. This is called a Gurney flap. As a matter of fact, tabs like the Gurney flap can be mounted on any element in the multi-element high lift systems to increase the loading of the main element and delay flow separation on the trailing edge. In the other way, high lift can be produced by the delay of stall which is seen in the phenomena of airfoils and wings oscillating in pitch and has a maximum angle of attack greater than the static stall angle. The mechanism of dynamic stall appears on helicopter rotor blades, rapid maneuver aircraft and wind turbines, and even insects. Dynamic stall was explored by McCroskey and Fisher [l]in an experimental investigation of a model rotor. The dynamic effects were verified to be the result of a vortex-dominated flow field. In this study, a dual time-step integration method with a preconditioned pseudo time derivative term is used. A RNG algebraic turbulence model is used to evaluate turbulence assumption. Numerical validation is performed on the evaluation of a NACA 4412 airfoil with a Gurney flap. A dynamic stall process associated with a NACA 0012 airfoil, oscillating in pitch at low reduced frequency and high Reynolds number in the experiments of McCroskey et al is also implemented in the numerical simulation. Summary: The two-dimensional time-dependent Navier-Stoke equations including a preconditioned pseudo time derivative term in conservative non dimensional form is

where,Q,=[p,u,v,t], Q=[ , u, v, e], F=[ u, uu+p, uv, u(e +p)] and G = [ v, uv, vv + p, v(e +p)]. Fv and Gv, are the viscous flux vectors.

where , u, v, p, a, e and H are the density, velocity components along x-, y- directions, static pressure, sound of speed and enthalpy, respectively. is assumed as the local velocity. Real time and pseudo time are denoted by t and , respectively. In the RNG based algebraic turbulence model, the integral scale is assumed to be proportional to the boundary layer thickness 6, and the eddy viscosity is obtained from

where v = vl + vt , subscripts I and t refer to the turbulent viscosity and laminar viscosity, respectively. The parameter a=0.0192, H is the Heaviside step function and is the dissipation rate. First of all, the NACA 4412 airfoil with different sizes of Gurney flaps such as one, two and three percents of airfoil chord length are taken as examples for the validation of the computation. Simulated results are compared with the predicted lift coefficients obtained by use of the INS2D code that has incorporated with the Baldwin-Barth turbulence model . Numerical prediction is performed at Mach number 0.085, Reynolds number 1.64 million, and zero angle of attack. From the Mach number contour distribution for a NACA 4412 airfoil with a Gurney flap of twopercent airfoil chord as shown in Figure I, a flow separation zone is formed at the backward side of the perpendicular plate, and a pair of contra rotating vortex can be found as revealed in Figure 2. From Figure 3, we can see that the predicted lift coefficients of the NACA 4412 airfoil with different sizes of Gurney flaps are close to the results of the Baldwin-Barth turbulence model. Obviously, the lift is increased as the variation of the length of the Gurney flap from 1 percent to 3 percent of airfoil chord length. It should be noted that the Gurney flap reduces pressure behind the flat plate, causing strong downward momentum of flow at the suction side of the airfoil. It is also found that the adverse pressure gradient around the trailing edge of airfoil is overcome, and a pressure recover region is formed at the suction side of the trailing edge. Large pressure difference between the upper and lower surface of the airfoil, caused by the effect of Gurney flap, results in increased lift.

Conclusion: A dual-time integration method with Roe s scheme is applied to predict aerodynamic characteristics of the NACA 4412 airfoil and an oscillating NACA 0012 airfoil. For the NACA 4412 airfoil, the predicted lift coefficients of the NACA 4412 airfoil with different sizes of Gurney flaps are close to the validated data. The production of lift is increased as the length of the Gurney flap changes from 1 percent to 3 percent of airfoil chord length. In the dynamic stall problems of an oscillating subsonic airfoil, the corresponding boundary-layer behavior for dynamic stall is observed in the current numerical simulation. The unsteady lift is increased until the pitching motion reaches near the maximum angle of attack. At that time, a strong negative pitching moment is induced. Numerical estimates of unsteady lift coefficient distributions are satisfactory compared with experimental data. However, the prediction of moment coefficient distributions needs further improvement.

Das könnte Ihnen auch gefallen