You are on page 1of 17

1

Airfoil Final Report


ME 332- Introduction to Fluid Mechanics Tuesdays at 2:40 pm under Aiswarya Due 4/19/13 Megan Dmello James Miller Brooke Peruski Scotty Schimp Miles Turrell

Abstract
This lab project involved the performance of 3 experiments on the symmetric NACA 0012 airfoil and the asymmetric NACA 4412 airfoil. The three experiments that were performed included the determination of the effect of varying the free stream velocity for a constant angle of attack on the drag and lift characteristics of the airfoil, determination of the effect of varying the angle of attack for a constant free stream velocity on the drag and lift characteristics of the airfoil, and the measurement of the pressure distribution on the airfoil for varying angles of attack. The results indicate that the asymmetric airfoil demonstrates far greater lift forces than the symmetric airfoil, particularly for angles of attack at its stalling point, which is at an angle of attack of 13 degrees. If the NACA 4412 asymmetric airfoil were to be used as the wing geometry of an airplane, care would need to be taken to ensure that the wings had a positive angle of attack of 13 degrees relative to the velocity vector of the airplane in motion.

Table of Contents
Introduction Page 4 Experimental Setup Page 6 Results and Discussion Page 10 Conclusion Page 16 List of Works Cited Page 17

Introduction
The purpose of this experiment is to determine the relationship between the angle of attack of the NACA 0012 and the NACA 4412 airfoils and the lift and drag characteristics of those airfoils. Additionally, the effect of an altered angle of attack on the static pressure profile of each airfoil for a constant velocity is determined. These experiments have a large practical significance due to the ubiquitous use of airfoils within the aerospace industry, particularly for the design of aircraft wings, helicopter blades, and other structures that produce lift. The extreme importance of airfoils lies in their low drag and their ability to produce lift provided the proper angle of attack and the ideal volume distribution. This property, known as camber, corresponds to the degree by which the volume of the airfoil is distributed above its centerline. An airfoil with a high percent camber appears to be top heavy. Increasing the camber increases the lift of the airfoil to an extent. Similarly, increasing the angle of attack, the amount of upward tilt the front of the airfoil has relative to the flow, will also increase the lift to an extent. By determining dimensionless relationships between the influence of camber and angle of attack on lift and drag, insight can be made towards designing better aircrafts. The NACA airfoils were airfoil shapes devised by the National Advisory Committee for Aeronautics (NACA). The digits that describe the airfoils are part of a 4-part code used to recreate the airfoil geometry (Heffley). If the 4 digits for the airfoil are known in conjunction with the airfoil chord, or length, the airfoil can be recreated using a series of equations. The first digit describes the maximum camber as a percentage of the airfoil length. So the NACA 4412 airfoil has 4 percent camber, indicating that it is asymmetric, while the NACA 0012 airfoil has 0 percent camber, indicating that it is symmetric. The second digit describes the location of the maximum camber along the chord line from the front of the airfoil in tenths of a chord. An airfoil with 0 percent camber would automatically have a second digit of zero. The max camber is not the maximum thickness. Rather, it is the maximum

5 amount that the airfoil line of center of mass shifts from the chord line due to camber. So the NACA 4412 airfoil has its location of maximum camber 40% of its total length from the nose of the airfoil. The final two digits describe the maximum thickness of the airfoil in percent chord. So the NACA 0012 airfoil maximum thickness is 12 percent of its length and the NACA 4412 airfoil also has a maximum thickness that is 12 percent of its length. Figure 1 shows the elements of the previously discussed airfoil.

Figure 1: Airfoil Diagram (Heffley)

In order to determine the relationship between airfoil angle of attack and the lift and drag of the airfoil, the lift and drag coefficients are plotted against the airfoil angle of attack. The lift and drag coefficient formulas are obtained from the Conservation of Linear Momentum Background Information lab document of the fluids lab. The drag coefficient, CD, is displayed in Equation 1. The lift coefficient, CL, is displayed in Equation 2. The Greek symbol Rho corresponds to the density of the

6 air in the wind tunnel. V corresponds to the free stream velocity within the wind tunnel. The symbol A corresponds to the cross sectional area of the wind tunnel central chamber. The symbol FD corresponds to the drag force in Newtons, and FL corresponds to the lift force in Newtons.

CD =

FD

1 rV 2 A 2 FL CL = 1 rV 2 A 2

(Equation 1)

(Equation 2)

Experimental Setup
The open loop wind tunnel is used for the airfoil lift and drag experiments. A pressure tap is located at the plenum of the wind tunnel to measure atmospheric pressure. A second pressure tap is located just in front of the airfoil to measure static pressure of the free stream. From the difference of these two values, the dynamic pressure, and thus the free stream velocity, can be computed using Bernoullis equation. The NACA 0012 and NACA 4412 airfoils are mounted on a force transducer stand by means of two bolts. The force transducer stand is located within the central chamber of the wind tunnel. The first bolt is used to mount the airfoil to the stand such that it can rotate about that point. A secondary screw is used to lock the angle of attack of the airfoil in place by means of friction between the screw and the force transducer stand. By slightly unscrewing the screw, the angle of attack of the airfoil can be manually changed. The screw is then tightened to lock the airfoil angle in place. The airfoil-stand assembly is adjustable by removing a wind tunnel cap adjacent to one side of the airfoil and reaching into the central chamber. The wind tunnel cap has an outline of the corresponding airfoil such that it can be lined up with the airfoil. There are a series of angle measurements around the perimeter of the cap such that the angle of attack of the airfoil can be measured after it is adjusted.

7 However, there is no method for determining the angle of attack while simultaneously moving the airfoil, so a great deal of estimation is required in the adjustment of the airfoil orientation to the desired angle. When the wind tunnel is turned on, transducers corresponding to the horizontal and vertical reaction forces of the stand output voltages to two multimeters. The multimeter readings are then recorded by the Labview software and converted to force readings by means of the mp constant value. The drag force and the lift force are plotted on their own plots in relation to inputted angle of attack, based on the measured angle of attack of the airfoil. The lift and drag forces in relation to the angle of attack measured in the Labview software are recorded into a separate excel document. The excel document contains a column with the desired angle of attack, the actual angle of attack, the pressure differential, and the mp values corresponding to the static and atmospheric pressure transducers. The second experiment involves changing the free stream velocity of the wind tunnel for a given angle of attack of each airfoil. For this experiment, the angle of attack is chosen to be near the stalling point. This is the point where the lift force starts to decrease for an increasing angle of attack. The angle of attack is kept constant throughout these experiments. In much the same manner as the experiment discussed in the previous paragraph, the horizontal and vertical reaction forces are recorded and plotted on Labview. A symmetric airfoil with a zero degree angle of attack should theoretically have no lift force. However, experimental results indicate that this assumption is not entirely correct. Therefore, the lift forces for all trials are normalized using the lift force on the symmetric (NACA 0012) airfoil at 0 degrees. Therefore, the normalized lift force on a symmetric airfoil at zero angle of attack is zero for all velocities. The asymmetric airfoil theoretically experiences lift at a zero degree angle of attack, because due to its geometry the flow moves faster along its top than its underside. However, in order to account for uncontrollable lift additions from the equipment and the experimental setup, the lift force on the symmetric airfoil at zero degree angle of attack is subtracted from all asymmetric airfoil lift forces.

8 The third experiment determines how the static pressure distribution about the NACA 0012 and NACA 4412 is influenced by their angle of attack relative to the wind tunnel free stream. The free stream is set to a constant velocity throughout each trial. Each airfoil has a series of static pressure taps about their perimeter, and these pressure tap readings are converted into voltage readings using a Validyne pressure transducer. In order to conserve experiment resources, a Scannivalve motorized scanning system is implemented to measure each successive pressure tap reading with a single pressure transducer. In this way, the multiple pressure taps around the airfoil do not each require their own pressure transducer. The voltage readings are then recorded, converted by means of the mp constant into their corresponding pressure values, and plotted using the Labview software. Figure 1 below shows the orientation of the pressure taps on the NACA 0012 symmetric airfoil. The arrow at the top of the airfoil shows how consecutive pressure ports are placed on the airfoil, from 1 representing the first pressure port to 18 representing the final pressure port. The pressure port locations of the NACA 4412 airfoil can be similarly determined based on Figure 3 below. The location of the pressure ports has an enormous influence on the graph layout and the interpretation in the results section.

Figure 2: NACA 0012 Airfoil pressure tap layout

Figure 3: NACA 4412 Airfoil

In order to compare the pressure distribution results from trial to trial, the surface pressure coefficient, Cp, is used. The surface pressure coefficient is the difference of the static pressure at the surface of the airfoil and the pressure of the free stream, divided by the dynamic pressure of the free stream.

Cp =

ps - p 1 2 rU 2

(Equation 3)

The velocity of the free stream in the wind tunnel is non-dimensionalized using the famous Reynolds number. The experiment that observes the relationship between free stream velocity and lift and drag coefficients plots the coefficients against the Reynolds number, not the free stream velocity. This allows the underlying relationship between the velocity and pressure to be observed, without needing to take into account the effects of the viscosity of the fluid or the diameter of the wind tunnel. The added benefit to only plotting non-dimensionalized variables is that the results obtained in this report can be applied to experiments using different wind tunnels or different fluids.

Re =

rVD m

(Equation 4)

The Reynolds number is displayed above. Rho corresponds to the density of the fluid, D corresponds to the characteristic length of the structure that the fluid is in, V corresponds to the fluid velocity and mu corresponds to the dynamic viscosity of the fluid. In the case of an airfoil in moving flow, the characteristic length is the chord length of the airfoil.

10

Results and Discussion


The results for the trials involving varying the free stream velocity and observing the lift and drag coefficient values are depicted in Figure 4. Based on these results, it is clear that the lift coefficient of the asymmetric airfoil reaches a maximum for Reynolds numbers of approximately 2500. For lower free stream velocities, which correspond to lower Reynolds numbers, the lift coefficient is lower. For Reynolds numbers higher than 3000, the lift coefficient of the asymmetric airfoil steadily decreases. The lift coefficient differences for the symmetric airfoil for varying Reynolds numbers are much less pronounced than for the asymmetric airfoil. This is intuitively understandable because the velocities at the top and the bottom of the symmetric airfoil are theoretically equal, and thus the static pressures from the Bernoulli equation on the top and the bottom of the airfoil are equal. The drag coefficients on the airfoils are very close to zero and have a slight but certain correlation with Reynolds number. Intuitively, as the velocity of the fluid passing around the airfoils increases, the molecules making up the fluid strike against and tug on the molecules in the airfoil at a higher rate. This would cause an increasing drag coefficient. The asymmetric and the symmetric drag coefficients seem to have nearly identical values for a given Reynolds number, indicating that the percent camber does not have a large effect of the drag on the airfoil. The magnitude of the force outputted by the horizontal force transducer was negative. It was assumed that the negative values of drag, which otherwise increased with increasing fluid velocity, were due to the direction of positive magnitude for the horizontal force transducer. The force transducer defined a positive force to be a force pointing upstream, while the drag force is clearly a force pointing downstream. If this distinction is made, and the absolute value of the drag forces is used, the drag coefficient increases in an intuitive manner.

11

5 4.5 4 3.5 Coefficient Values 3 2.5 2 1.5 1 0.5 0 0

Lift and Drag Coefficients as a Function of Reynolds Number

Symmetric Drag Coefficient

Symmetric Lift Coefficient

Asymmetric Drag Coefficient

Asymmetric Lift Coefficient

5000 10000 Reynolds Number

15000

Figure 4: Lift and Drag Coefficients as a Function of Reynolds Number

Figure 5 below depicts the affect of the airfoil angle of attack on the lift and drag coefficients of the symmetric and asymmetric airfoil. The lift coefficient of the asymmetric and symmetric airfoils both increase with increasing angle of attack up to their stalling points, at which their lift coefficients decrease. The stalling point for the asymmetric airfoil occurs at approximately an angle of attack of 13 degrees while the stalling point of the symmetric airfoil occurs at approximately an angle of attack of 11 degrees. The asymmetric airfoil has a higher lift coefficient than the symmetric airfoil for all positive angles of attack considered. Additionally, the range of lift coefficients for the asymmetric airfoil is much larger than for the symmetric airfoil. It is important to notice that the graph below confirms that the asymmetric airfoil still experiences lift for a zero angle of attack, while the symmetric airfoil does not experience lift. The asymmetric airfoil requires a negative angle of attack of approximately 2 degrees in order for it to also experience no lift force.

12 The drag coefficients for the symmetric and asymmetric airfoil remain at similar values for angles of attack below 10 degrees. For angles of attack between 10 and 15 degrees, the drag coefficient of the symmetric airfoil increases to a higher plateau. For the later angles of attack from 15 to 20 degrees, the asymmetric drag coefficient increases to a higher value. This indicates that the optimal angle of attack for a NACA 4412 airfoil is approximately 13 degrees, because the airfoil would experience the maximum lift, and its drag would still be similar to the drag for lower angles of attack.

6 5 4 Coefficient Values 3 2 1 0 -10 -5 -1 -2 0

Angle of Attack Variation


Symmetric Drag Coefficient Symmetric Lift Coefficient Asymmetric Drag Coefficient 5 10 15 20 25 Asymmetric Lift Coefficient

Angle of Attack (Degrees)

Figure 5: Lift and Drag Coefficients as a Function of Angle of Attack

The next experiment involves the pressure distribution on the NACA 0012 and NACA 4412 airfoils. The following graphs show the effect of the pressure port location on the airfoil on the pressure coefficient. The quantity x/c is used to represent the pressure port location, where c is the chord length and x is the distance the pressure port is from the leading edge along the chord. The pressure distribution on the NACA 0012 airfoil is shown in Figure 6 below. The negative pressure coefficient shown in Figure 6 corresponds to a lower static pressure at the top of the airfoil than the bottom of the airfoil. The airfoil

13 thickness sharply increases near its leading edge, which corresponds to a sharp decrease in pressure coefficient. This is because the radius of curvature near the front of airfoil is the shortest, so a large suction force normal to the streamlines around the front of the airfoil is created.
Cp vs. Location for Various Angles of Attack for NACA 0012 Airfoil
1.5

0.5

0 0 -0.5 Cp 10 20 30 40 50 60 70 80 90 100 0 deg 5 deg 10 deg -1 13 deg

-1.5

-2

-2.5

x/c

Figure 6: Pressure Distribution as a function of angle of attack of NACA 0012 airfoil

The pressure distributions on the NACA 0012 airfoil and the NACA 4412 airfoil are similar in some respects, and different in others. Figure 7 below depicts the pressure distribution on the NACA 4412 airfoil. Note from Figure 3 in the Experimental Setup section of this report that the NACA 4412 airfoil has pressure taps both on its top and bottom. The location of the pressure ports on the airfoil were not known for the NACA 4412 airfoil, so instead of plotting x/c on the abscissa of Figure 7 in the same manner as Figure 6, the port index values were used. The NACA 4412 airfoil, unlike the NACA 0012 airfoil, is asymmetric, so the pressure distribution on its top and bottom is different. This is not the case for the NACA 0012 airfoil; since the pressure distributions on the top and bottom of the NACA 0012 airfoil are identical, the entire pressure distribution over the airfoil can be determined with pressure taps on one side. Similarly to the NACA 0012 airfoil, the

14 NACA 4412 airfoil also has its sharpest radius of curvature near the front and top of the airfoil. This causes a similar negative pressure coefficient for a variety of angles of attack in the same manner as for the NACA 0012 airfoil. The pressure coefficient steadily increases after the location of smallest radius up to the airfoils trailing edge. Based on the angle of attack, the pressure distribution on the bottom of the NACA 4412 airfoil either increases in pressure from the trailing edge to leading edge, or decreases in pressure. For positive coefficients of pressure from the pressure ports 12 to 18, a positive Cp corresponds to a series of force vectors pointing into the airfoil, that is, a pushing pressure. For negative coefficients of pressure over the same ports, the series of force vectors point away from the surface of the airfoil, that is, a vacuum pressure is present. For angles of attack of 5 degrees or more, the pressure pushes on the bottom of the airfoil in decreasing magnitude from the tip to tail. For all angles of attack, the pressure distribution has a similar absolute magnitude behavior over the bottom of the airfoil as the pressure distribution on the top of the airfoil. The difference occurs when the sign of the bottom pressure is considered. It is clear that in order to maximize the total lift force on the NACA 4412 airfoil, it is best to tilt the airfoil to the free stream such that it has a sufficiently large angle of attack. Note the relationship between the lift-reducing pressure distribution corresponding to a zero degree angle of attack in Figure 7 and the rapidly reducing lift forces measured for the NACA 4412 airfoil in Figure 5. It is clear from both figures that the total lift force on the NACA 4412 airfoil is highly dependent upon its angle of attack relative to the free stream.

15

Cp vs. Port Number for Different Angles of Attack NACA 4412


1.5 1 0.5 0 0 -0.5 Cp -1 -1.5 -2 -2.5 -3 2 4 6 8 10 12 14 16 18 20 0 deg 5 deg 10 deg 13 deg

Port Number

Figure 7: Pressure Distribution as a function of angle of attack of NACA 4412 airfoil

16

Conclusion
This lab project involved the performance of 3 lab experiments on the symmetric NACA 0012 airfoil and the asymmetric NACA 4412 airfoil. The three experiments that were performed included the determination of the effect of varying the free stream velocity for a constant angle of attack on the drag and lift characteristics of the airfoil, determination of the effect of varying the angle of attack for a constant free stream velocity on the drag and lift characteristics of the airfoil, and the measurement of the pressure distribution on the airfoil for varying angles of attack. The results indicate that the asymmetric airfoil demonstrates far greater lift forces than the symmetric airfoil, particularly for angles of attack at its stalling point, which is at an angle of attack of 13 degrees. If the NACA 4412 asymmetric airfoil were to be used as the wing geometry of an airplane, care would need to be taken to ensure that the wings had a positive angle of attack of 13 degrees relative to the velocity vector of the airplane in motion.

17 Airfoil Final Report List of References

Heffley, David. Aerodynamic Characteristics of a 4412 Airfoil. Baylor University 26 January 2007. http://www.baylor.edu/content/services/document.php/41147.pdf Kooshesfahani, Manoochehr. Fluid Mechanics Laboratory. Copyright 2006.

You might also like