Sunteți pe pagina 1din 18

AERODYNAMCIS Final Design Report Chapter Draft 1 Team 2 Jesse Jones Chris Selby AAE 451 Fall 2006

AERODYNAMICS CHAPTER Introduction The high speed flight mission requirement was considered to be of foremost importance by the aerodynamic team. Flying fast requires high thrust. However, the thrust is limited by the selection of the motor. The motor has to put as much power out as possible and still fit in the budget. With this budget constraint in mind, the aerodynamic team minimized the drag as much as they could so that the aircraft could fly as fast as possible. Another mission requirement considered to be important by the aerodynamic team was the minimum stall speed at level flight requirement. This directly put a constraint on the CL, max the aircraft must have. Later this CL, max constraint would play a key role in the decision to include or not include flaps. Next a discussion will be presented on the selection of the airfoil sections for the wing and tail. The selection process follows the scheme highlighted above. First, the wing and tail were designed for high speed. Then, the wing and tail were designed to meet the minimum stall speed requirement. Wing Airfoil Design The mission requirements for this design dictate conflicting constraints on the wing design. The high speed dash demands a low lift, minimal drag airfoil; the slow stall speed will require a large maximum lift coefficient; while the economic loiter and cruise require a high L/D value. After some preliminary mission analysis it became evident that the high speed dash design point would be the most constraining to the airfoil selection. Stated differently, a high speed wing (low lift, low drag) can use flaps to achieve higher lift coefficients for stall condition, but a high lift airfoil cant easily be modified to achieve minimal drag at small lift coefficients. Additionally, the loiter and economic cruise requirements will likely be irrelevant, as a power plant sized for the high speed mission should have more than enough energy to meet the endurance requirements of the loiter and economic cruise. Having set the high speed cruise condition as the primary design point, the task was to find an airfoil with minimal drag precisely in the range of Cl values needed for the high speed condition. The first step, then, becomes calculating the required 3-D lift coefficient at the design point and relating this to an airfoils 2-D lift coefficient. The required CL value at stall can be found using Eq. (1) seen below. 2

C L ,cruise =

W 1 2 Vcruise S 2

Equation 1: CL, cruise

In the above equation, weight, cruise velocity and wing area are all design values which were still fluctuating at the time of airfoil selection, so the preliminary design values had to be used as an estimate (S = 4.5 ft^2, W = 5.5 lbs, V = 92 ft/s). These preliminary criteria imply a lift coefficient of 0.124 at the design point. To convert the 2-D lift coefficient to 3-D lift coefficient, lifting line theory was used. While the general lifting line equations are quite complex to solve, the choice of a 0.45 taper ratio approximates an elliptic lift distribution and enables a much simpler solution to the lifting line equations. The fundamental lifting line equation is presented below as Eq. (2) (see Reference 4).
( y o ) 1 ( y o ) = +L =0 ( y o ) + V c( yo ) 4 V

( d dy ) dy o y) b 2
b 2

(y

Equation 2: Lifting Line Theory

Note that is the geometric angle of attack as a function of the location along the span, yo. =0 is the zero lift angle of attack as a function of yo. c is the chord length as a L function of yo. y is a variable of integration over the span. V is the free stream velocity; and finally, is the circulation around a wing section as a function of yo. The end goal is to obtain a relation between CL as a function of angle of Cl. To do this a relation between gamma and angle of attack must be obtained, which can be used in the Kutta-Joukowski theorem, lift equation given below as Eq. (3) (Reference. 4). Note that an expression for the lift coefficient can be derived from Eq. (3) as shown below.
L' = V Equation 3: Sectional Lift

Cl V c 2

Equation 4: Circulation

In general, the desired expression would be quite a complex expression. However, for elliptical lift distributions, the circulation can be expressed quite simply as Eq. (5) below (Reference. 4). Note that b in this equation represents the total span of the wing.
2y ( y ) = 1 o b
2

Equation 5: Elliptical Circulation Distribution

By substituting the elliptical circulation distribution expression (Eq. (5)) into the fundamental lifting line Eq. (2), it can be shown that the chord distribution for a wing with elliptical lift distribution is elliptical. So for a wing with elliptical chord distribution the lift coefficient can be calculated using the following.
L = V o
b 2 b 2

b 2y 1 dy = V o 4 b

Equation 6: Elliptic Lift Distribution

V o b4 b L CL = 1 2 = 1 2 = V S 2 V S 2V S 2
Equation 7: Elliptic Lift Coefficient

Substitute in Eq. (4)


CL =

b Cl V c = Cl 2V S 2 4

Equation 8: Elliptic Lift Coefficient

\Plugging the previously calculated, required CL=0.124 into the above equation yields a required Cl = 0.158, the target minimum drag lift coefficient for our airfoil. There are, seemingly, an endless number of airfoils to choose from; and there isnt time to analyze them all. In the end, several NACA 4-series airfoils were analyzed as they are a staple of aerodynamics and plenty of experimental data is available for them. Also, NACA 6-series airfoils were examined for similar reasons to the above but also due to their design for high speed / low drag. The Martin Hepperle series of airfoils were examined as likely candidates as many of them are designed for model pylon racer airplanes, which share the low Reynolds number high speed mission of this design. Finally, the team also made an effort to create an R.T. Jones type airfoil that would meet our design needs. The R.T. Jones airfoil design was conducted utilizing a code created by Dr. J. Sullivan, Professor, Purdue University. To ascertain the performance behavior of these various airfoils, an airfoil design tool called XFOIL (created by Mark Drela) was utilized. The viscous subroutine of XFOIL was utilized for these calculations with the following parameters.
Table 1: XFOIL viscous parameter identification
Reynolds Numer Mach Number Transition Criterion 500,000 0.15 e9

The XFOIL results for some of the more likely candidates are seen in Fig. (1) in Appendix A. As can be seen in this figure, there were several airfoils that met the design requirements. After discussion of the XFOIL results and prospective airfoils, it was decided that the NACA 1408 would be the airfoil design for the main wing. This decision was based on the 1408s drag bucket aligning most closely with the aircrafts high speed lift coefficient range, as well as the abundance of experimental data available for the airfoil. The experimental data was expected to be crucial in determining a dependable Cl, max value. Tail Airfoil Design Keeping with the theme of design for high speed cruise, the primary consideration for tail airfoil selection was to minimize drag. Some aircraft designs utilize a lifting horizontal tail section to more efficiently counter thrust and loading moments or to supplement the lift of the main wing. However for the sake of simplicity, as a result of the infant condition of the loading and thrust designs, and under the presumption that the loading and propulsion design would attempt to minimize the moments at the high speed cruise condition, the horizontal tail section was predetermined to be symmetric. As is usually the case, the vertical tail was also predetermined to be symmetric. One other significant constraint on the tail airfoil selection is that they had to be thick enough for structural and manufacturability purposes. The aerodynamics of drag would dictate an extremely thin airfoil (at zero angle of attack), but this would lead to a wing that was impossible to manufacture and/or structurally inadequate. Using rough geometric scaling, a preliminary chord estimate of six inches was agreed upon; and after consultation with the structures design group, a minimum thickness of 7% was established (implying a thickness of less than half an inch). From this point, the strategy was to analyze the drag performance of various symmetric airfoils. The obvious candidates for symmetric airfoils around 7% thicknesses were the NACA 0006, 0007, and 0008. The team also made an effort to create an R.T. Jones type airfoil that would meet our design needs. As stated in the previous section, the R.T. Jones airfoil design was conducted utilizing a code created by Dr. J. Sullivan, Professor, Purdue University. Drag behavior of the various candidate airfoils was again calculated using XFOIL. The viscous subroutine of XFOIL was utilized for these calculations with the parameters listed in Table (1) above. The calculations from XFOIL are plotted in the appendix as Fig. (2) and Fig. (3). Fig. (3) is simply a close-up at the smaller angles of attack. A first attempt was to make the 5

vertical tail and horizontal tails flat plates. Upon analyzing this configuration in XFOIL, the flat plate had enormous drag penalties compared to the other airfoils being tested as seen in Appendix A Fig. (4). The next consideration was discerning the likely angle of attack range of the tail wings at the high speed cruise condition. It was determined that the vertical tail should be relatively easy to attach to the aircraft at a fairly accurate zero angle of attack. Also, there should be no significant steady state forces or moments that would drive the vertical tail from zero degrees angle of attack (in fact, zero degrees angle of attack should represent a stable equilibrium for the vertical tail). This implies that the best airfoil for the vertical tail is the one with the least drag at very small angles of attack (less than 2). Fig. (3) shows that one of the Jones airfoils that the team designed best met this criterion. The R.T. Jones / Joukowski parameters that define this airfoil are presented below in Table (2)
Table 2: Vertical Tail Jones/Joukowski definition
Parameter xc yc xt yt Value -0.0617 0 1 0

The horizontal tail requires some alternative considerations. The first consideration is that from a practical standpoint, it will be difficult to mount the horizontal tail precisely at zero degrees angle of attack (due to the lack of a well defined longitudinal axis). Also, due to the likelihood of unbalanced wing and loading moments at the high speed cruise condition, the horizontal tail will probably be in equilibrium at some small angle of attack other than zero (but say, less than 5 degrees). For this reason it is desired that the horizontal tail section have a low drag coefficient over a range of angles of attack. Again, looking at the XFOIL results in the appendix, it is apparent that the NACA 0008 is the best choice for the horizontal tail. The horizontal tail span was set due to a rather unconventional requirement. Since the aircraft configuration utilizes a dual boom tail with the booms connecting to the main wing, the span of the tail affects the main wing. Specifically, it would be mechanically more complex (and likely cause more interference drag than added lift) to have flaperon sections inboard and outboard of the location of the boom mounts on the main wing. Also, since the main wing section has minimal camber, a large CL, max will be required of the flaps. Having a large tail span will limit the span of the flaps. On the other end, having a tail span that is relatively small will decrease the control authority of the horizontal tail and cause interference drag between the booms and the fuselage. As such, after consulting with the 6

dynamics and controls group and performing some preliminary calculations of tail sizing, the horizontal tail span was set at 18 inches. The dimensions of the vertical tail as well as the area (and implicitly the chord) of the horizontal tail were determined via dynamics and controls considerations discussed later. Cl, max and CL, max Estimation without Flaps The value of Cl, max can be found from experimental data for NACA airfoils at Reynolds numbers of 3e6, 6e6, and 9e6 (see Reference 3). However, the aircraft is not designed for these high Reynolds numbers so the values of Cl, max from the experimental data at these Reynolds numbers are not relevant. A viscous panel method program, XFOIL, is used to determine the aerodynamic data for an airfoil at a Reynolds number relevant to the design. To validate the results of XFOIL, the experimental data was compared to XFOIL data obtained at the same Reynolds number. Fig. (5) Fig. (7) in Appendix A show that XFOIL consistently stalls at an stall that is 25% bigger than the actual stall. Another way of looking at this correlation is that the actual stall is 80% of the XFOIL predicted stall. The XFOIL Cl value at this corrected stall is then taken to be the actual Cl, max. Refer to Fig. (8) for the predicted XFOIL Cl, max value for the NACA 1408 airfoil using the 80% stall rule. To get the 3-D CL, max coefficient the designers looked at equations and charts presented in Reference 1, Eq. (12.15). This equation is represented below as Eq. (9).
C L , max = 0.9C l , max cos 0.25 c Equation 9: CL, max estimation from Cl, max

For the wing, the aerodynamic and structures team decided to make the quarter chord angle 0 degrees for ease of representation of aerodynamic loading. This simplifies Eq. (9) so that 3D CL, max is 90% of 2-D Cl, max. Other methods of relating 3-D CL, max to 2-D Cl, max are presented in Reference 1 and when they are simplified for the low mach numbers this design operates at, the result that 3-D CL, max is approximately 90% of 2-D Cl, max still holds. Adding the effects of flaps increases the CL, max for the wing and will be discussed in detail next. CL, max Estimation with Flaps The design team chose to break down the CL, max of the aircraft into a flapped and unflapped CL, max. The unflapped portion of the wing would contribute to CL, max unflapped and the flapped portion of the wing would contribute to a CL, max flapped. The fuselage and tail surfaces were ignored because they were assumed to generate negligible lift. Elevator deflection was also ignored since this control surface would create more pitching moment 7

than necessary just to get a higher lift coefficient. The mathematical model used to represent CL, max is shown below in Eq. (10).
C L , max = C L , max
clean

S unflapped S

+ C L , max

flap

S flapped S

Equation 10: CL, max due to flaps

The above equation can be rearranged to look like Eq. (11).


C L ,max = C L , max
clean

S flapped S + C L , flap S S

Equation 11: CL, max due to CL, flap C Where, L , flap = C L , max
flap

C L , max

clean

. The maximum CL values (clean and flapped)

where obtained from XFOIL using the 80% stall rule and the 90% 2-D Cl, max rule both described in the previous section. Sflapped is the area of the wing the flaps affect. This depends on where the flaps are located and the CL, max due to flaps depends on the flapped percent chord of the wing (cf/c). The second of these two (cf/c) was optimized first to give the maximum CL for a flap deflection of 30. A maximum flap deflection angle of 30 was chosen so that the flaps could be deployed an additional 15 individually to act as ailerons giving the airplane more controllability in case the . XFOIL was used to analyze different flap (cf/c) configurations until one was found with the need CL, flap to meet the CL, max requirement. Increasing the CL, max any more than needed would result in larger and larger percent chord flaps which are unwanted. The larger percent chord flaps means that the servos have to be bigger so that they could move the control surface. Bigger servos mean more weight, and more weight means that the aircraft will not fly as fast. Design Parameters The most important aerodynamic parameters are those that have to be determined before the aerodynamic design can begin and are based primarily on historical data or a designers initial guess. Generally these important design variables are varied at a later stage in the design to determine an optimum value for each. Some of the most important aerodynamic design variables in this airplane design are: wing aspect ratio (AR), wing taper ratio (), maximum flight speed (Vmax), wing planform (S), fuselage diameter (dfuse), and fuselage length (lfuse). There are more parameters, but the ones presently listed effect the end results much more significantly and so they will be discussed in further detail below. Aspect Ratio 8

The wing aspect ratio was initially chosen to be 7.6 upon the recommendation from Reference 1, Table 4.1. An aspect ratio of 7.6 could be found in most general aviation single engine aircraft. However, upon conducting a trade study to see how drag is affected by changing the aspect ratio, it was found that a lower aspect ratio of 5 is more desirable. This new aspect ratio of 5 gives a lower parasite drag and thus lower thrust loss and greater speed. Taper Ratio A taper ratio of 0.45 was recommended by Reference 1, in Chapter 4. This taper ratio is shown through the use of lifting line theory to most accurately produce an elliptical lift distribution along the wing. Prandtl proved in the early 20th century that an elliptical loading produces the least amount of induced drag. Reducing the drag on the aircraft by using a 0.45 taper ratio will decrease the thrust required for all speeds, allowing a greater maximum speed. Flight Velocity The maximum flight speed was initially chosen to be 150 ft/sec. Upon further analysis from the propulsion team, this value was then lowered to 92 ft/sec because of budget constraints. The flight velocity affects the aerodynamic performance through dynamic pressure. Aerodynamic parameters such as CL, max and CL, min scale with dynamic pressure. These parameters, among others, are an integral part of calculating the different aerodynamic variables. So, changing the flight speed changes most of the variables in most of the calculations. The end result is that as the flight speed decreases, the parasite drag increases. A brief explanation for this is the decrease in speed decreases the Reynolds number which in turn increases the skin friction coefficient. The increase in skin friction then causes an increase in the parasite drag. Wing Planform The wing planform was determined from initial sizing. In the initial sizing of the aircraft, a wing loading was determined. An overall weight of the aircraft was then guessed and from the wing loading, a planform area of the wing was found. The final sizing of the planform was calculated from Eq. (12) below. The final planform is found based on updated weight estimates, the initial value of CL, max that the wing and flap system could get, and the 9

stall speed mission requirement. The method used to find the CL, max was outlined earlier in the CL, max Estimation with Flaps section of this document.
S= W 1 V 2 C L , max 2

Equation 12: Modified CL, max equation solved for wing planform

Fuselage Diameter/Length The fuselage diameter has a minimum value set forth by the payload dimensions and material thicknesses. For parasitic drag issues, the diameter of the fuselage is desired to be as small as possible. The length of the fuselage is also of great importance in the calculation of parasitic drag and must be as small as possible, but big enough to fit the payload and the avionics of the aircraft. To reduce the drag caused by the fuselage, the wing is blended into the fuselage by the use of fillets. This decreases the interference drag caused by the wing and fuselage but does not eliminate it. Another way to decrease the interference drag due to the fuselage is to keep the shape of the fuselage as cylindrical as can be constructed. Doing this would decrease the form drag because sharp corners tend to generate more drag than smooth corners. The method used to estimate the parasite drag is described below. Drag Component Buildup Method A drag build-up method was used to compute the parasite drag. This method is outlined in Reference 1, Chapter 12 and the paper in Reference 2. This method estimates the subsonic parasite drag of each component modeling each component with flat-plate skinfriction drag coefficient and a component form factor. The form factor estimates the pressure drag due to viscous separation. An interference effects is also used for components such as fuselage and wing interference. Eq. (13) below is taken from Reference 1, Eq. (12.24) and shows the calculation for the estimation of CD0.
C D0 =

(C

f ,c

FF c Qc S wet ,c ) S ref

Equation 13: Drag Component Buildup Method

The skin friction coefficient for each component (Cf, c), is modeled after a flat plate. The flat-plate skin friction coefficients are a function of the local Reynolds number. The turbulent skin friction coefficients are modeled in Reference 2, Fig. (3) This figure is reproduced in Appendix A Fig. (9). Using the characteristic length of each component, a 10

local Reynolds number can be found and the component skin friction coefficient can be computed. Then the component form factor is applied along with any interference effects. Other terms not shown in Eq. (13) are the various fudge factors described in Reference 1, Chapter 12. For example, Raymer suggests a form factor about 10% higher than the one described in his text for a tail surface with a hinged rudder. So, the tail surface would have and extra 1.10 multiplying the other drag component terms (Cf, c, FFc, Qc, and Swet, c). Once the individual drag components are computed, they are summed up and divided by the reference area which is taken to be the wing planform area.

11

REFERENCES 1. Author, D. Raymer. (2006). Aircraft Design: A Conceptual Approach. (4th Ed.). Reston, Virginia: AIAA. 2. Dr. L. M. Nicolai. (2002). Estimating R/C Model Aerodynamics and Performance. [Electronic version]. Lockheed Martin Aeronautical Co.. 3. Abbot, Ira H. and Von Doenhoff, Albert E. (1975). Theory of Wing Sections. New York, NY. Rupple Publications

12

APPENDIX A Figure 1 XFOIL Drag Polar for Various Wing Airfoil Sections [Insert plot later] Figure 2 XFOIL tail airfoil section data [0-5 range] Figure 3 XFOIL tail airfoil section data [0-3 range] Figure 4 XFOIL tail airfoil section data (flat plate effect) Figure 5 XFOIL vs. Experimental Cl- curve at Re = 3e6 Figure 6 XFOIL vs. Experimental Cl- curve at Re = 6e6 Figure 7 XFOIL vs. Experimental Cl- curve at Re = 9e6 Figure 8 XFOIL predicted Cl, max of NACA 1408 airfoil Figure 9 Skin Friction Coefficient vs. Reynolds Number

13

Vario
0.013

0.012

0.011
Figure 2: XFOIL tail airfoil section data [0-5 range]

0.01

0.009 Cd 0.008 0.007 0.006

14

Vario
0.008

0.007 3: XFOIL tail airfoil section data [0-3 range] Figure

Vario
0.025 0.006

Cd

0.02
Figure 0.005 4: XFOIL tail airfoil section data (flat plate effect)

15

1.5
Figure 5: XFOIL vs. Experimental Cl- curve at Re = 3e6

Cl

0.5

1.5 -10 Figure 6: XFOIL vs. Experimental C - curve at Re = 6e6 -2 -8 -6 -4


l

0 0
16

1.5
Figure 7: XFOIL vs. Experimental Cl- curve at Re = 9e6

Cl

1.2

0.5

Cl
0
17

1 -10 0.8
Figure 8: XFOIL predicted Cl, max of NACA 1408 airfoil

0 -5

Figure 9: Skin Friction Coefficient vs. Reynolds Number

18

S-ar putea să vă placă și