MARS UNMANNED AIRCRAFT 2003-2004 ODYSSEUS TEAM CORNELL UNIVERSITY ADVISOR PROFESSOR MICHEL Y. LOUGE TEAM LEADER ALEXANDER CHEFF HALTERMAN DATE APRIL 1, 2004 THE TEAM: Team Members: Alicia Billington Emmanuel Franjul Jian Gong Alexander Halterman Yen-Khai Lee Jeremy Nersasian Cem Ozkaynak Jing Pei Mikiko Ujihara arb351 ef35 jg253 ach22 yl245 jbn5 co37 jp292 mu23 BEE MAE ECE MAE ECE MAE ECE MAE MAE (2006) (2005) (2004) (MEng 2004) (2004) (MEng 2004)2 (2005) (2004) (2004) Advisor: Professor Michel Louge 1 2 myl3 Student ID numbers are the students e-mail address (ID#@cornell.edu i.e. arb35@cornell.edu) Ceased doing work after December 2003 due to January 2004 graduation Abstract As space exploration progresses, Mars gains more focus as the next frontier in human exploration. Manned missions to Mars have been discussed and planned to a certain degree. However, before humans can set foot on Mars, a wealth of information about Martian conditions will need to be provided by satellites, unmanned vehicles and a myriad of other data collection instruments. The goal of the Odysseus team is to design an unmanned aerial vehicle (UAV) for flight in the lower regions of the Martian atmosphere. Such a vehicle would collect specific, high resolution topographic data for speculated landing sites. The data returned by this aircraft will be of the utmost importance to the success of any Mars landing mission. This paper focuses on the aerodynamics, propulsion, structures and electrical systems of an unmanned aerial vehicle for flight on Mars. The Martian environment, as well as the mass sensitive nature of current space exploration, present a set of conditions by which an aircraft must be designed. Such a design must optimize mass, volume, flight time, power, and instrumentation in order to create an aircraft that can be sent to Mars with existing spacecraft while satisfying its exploratory goals once it reaches Mars. A UAV design for Mars must incorporate the aerodynamic properties for sustained flight in a low density atmosphere, propulsion in an atmosphere lacking sufficient oxygen for combustion, structural integrity with minimal mass and electrical controls for unmanned flight. Our UAV design overcomes these daunting constraints and provides a robust platform for reconnaissance of Mars. The final UAV design consists of a 10.8 kg aircraft with fuselage length of 2.1 meters, maximum fuselage diameter of 0.25 meters, and a wingspan of 2.078 meters. For propulsion we have chosen a single 2.27meter diameter three-bladed propeller mounted aft of an inverted V-tail. The propeller motor as well as the topographical and control instrumentation aboard the UAV is powered by multiple lithium-ion SPE batteries. The data produced during flight will be continuously transmitted to satellites orbiting Mars that relay the signals to Earth. Flight control and navigation is accomplished through feedback from on-board sensors that detect acceleration, pitch and roll. This design provides a flight time of 2.3 hours at a cruising speed and altitude of 130 m/s and 500 meters respectively. Further details of our design choices and potential alternatives are discussed within the following pages. fuselage designs. Electrical evaluated the capabilities of various electrical components and communication network designs. Through a series of presentations to Professor Michel Louge that focused on our conceptual, preliminary, and final design, we narrowed down the initial trade-studies to a single UAV design optimized for the Martian environment. Our design process focuses only on the technical aspects of flight on Mars, from the time the UAV is deployed through its expected life-cycle. Other aspects of the mission, such as launch costs and procedures, Earth-to-Mars transit routes, aero braking techniques upon reaching Mars, and atmospheric deployment feasibilities were not considered in our design. Additionally, cost, environmental impact, political motivation and human safety were not prevalent issues for us; however, technological and mission feasibility were carefully evaluated throughout the design process. Our final design only employs technologies that are currently available or on the horizon while acknowledging the many aerodynamic, structural, propulsion, and electrical challenges of deploying a fully-autonomous UAV on Mars for long periods of time. Introduction Mars is indisputably the centerpiece of current space exploration with both the scientific community’s space exploration efforts and the general public’s interest focused on recent investigative missions to the Red Planet. In the spirit of human exploration the Odysseus Team is designing an Unmanned Aerial Vehicle to fly on Mars as part of the Revolutionary Vehicles: Concepts and Systems University Student Competition for 2004 sponsored by NASA. A UAV serves as a reconnaissance platform for future manned missions to Mars. The aircraft will carry topographical and imaging instrumentation to produce valuable data regarding the Martian environment. The design of a UAV must optimize mass, volume, flight time, power, and instrumentation in order to create an aircraft that can be sent to Mars with existing spacecraft while satisfying its exploratory goals once it reaches Mars. Such a design for Mars must incorporate the aerodynamic properties for sustained flight in a low density atmosphere, propulsion in an atmosphere lacking sufficient oxygen for combustion, structural integrity with minimal mass and electrical controls for unmanned flight. Our UAV design overcomes these daunting constraints and provides a robust platform for reconnaissance of Mars. Atmospheric Constants T0 (K) 223 General Design Methodology: The design process began with the high-level abstract evaluation of the various objectives and constraints. Members of the team met twice a week to develop the mission profile, such as the scientific motivations behind a high-resolution topographical map of the Martian surface and atmospheric profile, both of which are not currently possible using existing satellites or landers. Following the mission profile, we consolidated the various engineering aspects – aerodynamics, structures, propulsion, electrical systems – into a systems-level hierarchy of the conceptual UAV design. Each level of the hierarchy takes into consideration safety and technological feasibility in order to identify critical paths and achieve our stated mission objectives. For instance, much of the fuselage design relied on the availability of a light-weight electrical engine to provide thrust needed to stay aloft in the low-pressure environment. As such, several interdependencies were stated early on and constantly revised as we progressed. Once the systems-level picture was developed, each sub-team developed its own set of trade-studies and constraints. Aerodynamics compared many airfoil and wing designs using Matlab and Excel. Propulsion compared the feasibility of propellers, chemical rockets, and jet engines given the atmospheric constraints on Mars. Structures required light-weight but durable P0(Pa) 700 (m-1 ) 9.00E-05 R 192.1 1.289 Table 1: Martian Atmospheric Constants Martian Atmosphere and Environment: Designing an aerial vehicle requires knowledge of the environmental conditions the craft will be flying in. One must know temperature, pressure, density and viscosity as functions of surface conditions and altitude. We have equations published by NASA that give acceptable and reasonable fits for atmospheric data. The equations are valid below 7000 Meters, which is above our cruise altitude, in an effort to capitalize on the largest density possible, these equations work well. T T0 .000998 * h (1) P P0 e h (2) Equations ( 1 ) and ( 2 ) give the temperature and pressure profiles respectively based on surface values. Where is a constant (9e-5m-1) and h is altitude in meters. From temperature and pressure we extrapolate data for density and speed of sound using known values and laws. The ideal gas law ( 3 ) gives us density as a function of temperature, pressure and the gas constant R, which is 192.1 on Mars. Equation ( 4 ) 1 solves for viscosity as a function of temperature. Speed of sound, a, is found using equation ( 5 ) and known values such as the ratio of specific heats, and temperature. T pR therefore flight at the correct altitude becomes important. Thrust specific fuel consumption (TSFC), a relationship between distance and fuel required, is ultimately a function of air density and thus altitude is the determining factor for fuel efficiency. TSFC sets an altitude for efficient cruise flight, from which a wing loading can be chosen to attain cruise conditions at the desired altitude. By choosing a specific wing loading the designer can fix the wing area for a craft with a target weight. Our design did not have the luxury of using thrust specific fuel consumption to fix wing area. Our aircraft will be propeller driven and run off electrical energy. Since our energy source has no dependence on pressure or density, we can not set an optimum cruise altitude. So we begin the design process without a specified wing area. With this area we would have been able to find appropriate airfoils, find maximum CL/CD values and proceed to optimize the aspect ratio dependant on wing weight and induced drag from wing end conditions. Without it, wing area becomes another parameter we need to optimize. Next we begin our search for suitable airfoils. Since the Martian atmosphere is approximately one hundredth the density of Earth’s and the craft is small in comparison to commercial aircraft, Reynolds numbers will be very low. Traditionally, planes fly well into the turbulent boundary layer regime, with a Reynolds number on an order of 106. With the conditions we are given, Reynolds number values will be between 40,000 and 80,000, with 80,000 being an extreme value that is unlikely. Dealing with such low Reynolds numbers poses a problem; boundary layers are largely laminar, which are notorious for flow separation due to low inertial forces. A craft flying in a laminar regime must utilize an airfoil designed specifically for low Reynolds flows. Traditional airfoil shapes are designed for turbulent conditions and will not suffice in the Martian atmosphere. On the advice of Professor David Caughey of Cornell University, we considered research done by Professor Michael Selig of the University of Illinois at Urbana-Champaign. Professor Selig has done a remarkable amount of work with low Reynolds airfoils and has a wealth of data available, including lift and drag coefficients at various angles of attack, as well as coordinates that can be used to numerically generate airfoils. With his data, we proceed with an airfoil design. We use Matlab and Excel to search through roughly 1500 of Professor Selig’s airfoils to find those most suitable for our applications. Rough calculations show that for velocities limited to Mach 0.6 and Reynolds numbers between 40,000 and 65,000, lift coefficients are best chosen to be 0.3-0.6. This is a relatively low value, but reasonable for this particular application. With low Reynolds numbers, due to the (3) 408.17 10 10 T 120 36.592T 1.5 asonic RT (4) (5) is usually seen to take the value 1.4 because that is its number for the air on Earth. is 1.289 on Mars due to the abundance of CO2 (over 95%) that makes up the Martian atmosphere. This lower value, along with the lower ambient temperature, causes a lower speed of sound, approximately 75% of that on Earth. This means that speeds need to be further limited to avoid sonic conditions. Other considerations on Mars include the rampant dust storms that can spring up unexpectedly and make flight very difficult. These dust storms are seasonal, allowing a wise launch and flight time to reliably eliminate this potentially devastating threat. Cruise Conditions h (m) 500 T (K) 222.5 P (Pa) 669.2 (kg/m3) 1.57E-02 1.45E-05 A (m/s) 234.72 Table 2: Cruise Velocity Atmospheric Conditions Aerodynamics Design Introduction As stated previously, the design of an aerial vehicle for Mars is a tricky process due to the planet’s harsh environmental conditions. Normally one begins an aircraft design process by defining take-off and landing scenarios, as well as approach to cruise altitude, but the fact that this vehicle will be launched from orbit makes this unnecessary. Our first step is to determine the wing loading necessary for the most efficient flight in cruise. Wing loading is the force per unit area on the wing during steady state flight conditions, which is important to determine for two reasons. First, it fixes the area of the wing. Knowing the wing loading simplifies the minimum drag analysis by tying the wingspan and mean aerodynamic chord (M.A.C.) together. Optimization then becomes a question of choosing the correct aspect ratio and taper ratio. The second reason involves the fuel efficiency; aircraft designed with combustion powered engines are sensitive to atmospheric changes; 2 need for a large chord and a small wing area, a lift coefficient that is too high would tend to limit aspect ratios. Induced drag becomes overwhelming in this situation, causing inefficient flight. With a rough value for the lift coefficient, we use Matlab to inspect different airfoils for high lift-todrag ratios in the correct regime. Fifteen airfoils are selected that have good characteristics around the aforementioned CL values. Finding a high ratio of CL/CD for 2D data does not guarantee a good airfoil. Rather, Reynolds number, maximum thickness-tochord ratio and actual lift coefficient are also important. Since the relationships are complicated and hard to judge by inspection we input the potential airfoils into Excel solver to find the optimum geometry and minimum drag for each airfoil. This completes the initial wing design. All that remains is to select the proper sweep angle that approximates an elliptical lift distribution. This can be done after the geometry is largely set, then optimized a second time to come up with the most efficient wing possible. equations, described in the previous Martian Atmosphere and Environment section of the report, that provide a rough model of the Martian atmosphere. We also assume the Selig airfoil data to be correct for all of the airfoils he tested. His tests were done in a 3 foot wind tunnel using a rectangular wing with a 33.375 inch span and a 12 inch chord. These dimensions suggest Selig’s decision to minimize 3D aerodynamic effects and that the data collected was analogous to that of a 2D airfoil. The fact that the span of the airfoil was so large in comparison to the tunnel strongly supports this theory. If this was not the case, blockage effects would be a large factor, and the tests would be inaccurate. Efforts for backing up this assumption are outlined later. Finally we assume that the wing weight estimation we use is accurate. Equation ( 8 ), later in the paper, uses a series of constants, as well as geometric and dynamic conditions, to make an estimate for wing weight given conditions. We use this data in optimizations to limit span. The values from this equation agree with the values the structures team obtain using the software Pro-Engineer and a suitable material. This appears to be a valid assumption. Design Assumptions: Before we begin the design process, we need to make certain assumptions to determine the optimal wing structure. These assumptions are made to account for the fact that we are not in possession of accurate data for every scope of our design process. If this design is to be finalized for physical flight on Mars, accurate weather data and atmospheric gradients need to be obtained to verify or refute our current calculations. This kind of data collection is out of the scope of this project and therefore we assume that our atmospheric representation is correct. We have a list of Lift Coefficient (CL) selection: We begin the wing design by finding a lift coefficient. Since the airfoils we are dealing with are for laminar flow, the drag data is very erratic, making it nearly impossible to find a valid curve that fits the data. Although the lift curve slopes from the data are very close to linear, the fact that drag is so far off makes curve fitting to find continuous points virtually impossible. Generally speaking, drag data from airfoils Figure 2: Drag vs. a For a Low Reynolds Airfoil Figure 1: Drag vs. a for a Standard Airfoil 3 C L vs. Reynolds Number will follow a parabolic curve (see Figure 1) in the region of the drag bucket. This allows you to fit a second order polynomial to the data, and find values of drag at continuous points on the curve. Figure 2 shows an example of drag data from one of the Selig airfoils we are considering. It can be seen that in low Reynolds airfoils laminar bubbles and possibly hysteresis in the switching from laminar to turbulent boundary layer conditions result in erratic data. A parabolic curve of the form C D C D kCL2 , where k is an arbitrary 0.6 0.55 0.5 CL 0.45 0.4 0 constant, can give an accurate fit to the drag data in Figure 1. Using this parabolic fit and the easily obtained lift curve slope, we can find CL and CD values for any angle of attack, given a wing planform and an aspect ratio. Since the data does not yield a valid curve fit, we are forced to use the discrete values that are provided with the Selig data. The best option for the laminar regime is to find a wing that is suitable for our purpose and test it for numerous angles of attack and use the data acquired. This is a time consuming process and is unfeasible given the time and resources available, so our choice is to use discrete data in place of more expansive experimental data. With discrete data we cannot determine an exact lift coefficient, as the chance of finding a suitable airfoil with that specific data point is unlikely. Instead, we consider a range of lift coefficient values dependant on Reynolds number. As Reynolds number increases, with all other variables being held constant, velocity increases and results in a lower lift coefficient needed for the same net lift. The opposite is also true; a low Reynolds number has a lower velocity, requiring a higher lift coefficient. For the high end values of Reynolds number, around 65,000, a CL value of around 0.35 0.3 35000 45000 55000 Reynolds Number 65000 75000 Chart 1: Optimal CL Values Re 40000 45000 50000 55000 60000 65000 CL 0.555 0.521 0.468 0.453 0.403 0.359 Table 3: Optimal CL for a Given Re 0.35 is optimal. At the other extreme, for the low Reynolds number case, a value of approximately 0.55 is optimal. These optimized CL values are determined using an Excel spreadsheet to generate feasible planforms for a given Reynolds number. With constraints set by the user and a specific value for Reynolds number, the solver generates optimal planforms for that condition. Angle of attack is one of the constraints, so an optimal value for CL is found for each airfoil and entered Reynolds number. Since drag is related to velocity and we desire a low value for drag, a low Reynolds number will yield a lower drag. A CL of 0.55 is set as the design lift coefficient with a corresponding Reynolds number of 40,000. However, this estimation does not take into account base drag of different airfoils, which play a small part in the overall drag. Further work shows that this CL value is optimal. Computational analysis: Computational analysis makes up a large percentage of the work involved in obtaining a sound model for the airfoil characteristics. Sorting through 650 airfoils is a manageable task with Matlab analyzing each airfoil and linking it to an Excel spreadsheet. The initial stages involve writing code to take the data and put it in a user friendly form. The design of a graphical user interface (GUI), shown in Figure 3 further eases the process of airfoil selection. The user can browse various airfoils based on target Reynolds number and vary plots of data until desirable curves are found. Saving the layout allows us to return to configurations at a later point for further review. This allows for Figure 3: Matlab User Interface for Airfoil Analysis 4 standpoint, a large AR, between 7 and 10, is a well designed wing. However, a large span means larger bending moments in the wing structure, due to lift loads acting farther from the root of the wing, resulting in additional weight needed to withstand the increase in bending moment; something discussed in the Structures section. Taper ratio is the ratio of tip chord to root chord, in our case 0.2. A wing with a low taper ratio, referred to as “a highly tapered wing,” tends to have lower lift coefficients on the outer portion of the wing, as the downwash pattern changes, toward an elliptical lift distribution. Low taper ratio also results in larger chords and wing thickness inboard where the bending moments are the largest, moving the lift in towards the craft, reducing the aerodynamic bending moments. Both of these effects are favorable for wing structural weight. However, low taper ratio wings have a tendency to stall at the tip, which is prevented with wing twist. With the final structural and electrical weights we find that a slight increase in wing area, to produce more lift, is required for cruise flight. We decided that we will keep the span constant and obtain the additional required platform area by increasing the taper ratio. The final taper ratio is 0.35, which leads to a platform area of 0.6 m2, a new AR of 7.2 and an increased safety factor of greater than 1.2 Geometric twist is the equivalent of taking a straight wing, and applying a moment about its axis, causing the tip of the wing to be at a different angle than the root of the wing, in our case, negative three degrees. A positive value of twist refers to an increase in angle of attack along the span of the wing. A negative twist value, known as “washout,” greatly increases the stability of the craft by causing stall to occur at the root of the wing before occurring at the wing tip. When this happens, control can still be maintained due to the moment that can be generated from the tips of the wings to control the rolling motion of the craft. Design Constraints Re specified taper ratio >.2 mach number Ma > .55 wing span 1.5 < b <4 altitude h >400 1/4 chord sweep 0 < c¼ <10 Aspect Ratio Ar >3 Table 4: Design Constraints simultaneous generation of several acceptable planforms for various airfoils. The ‘Target Re’ field allows the user to search for the airfoils closest to the desired Reynolds number. The fields that follow are outputted data relevant to the current airfoil. This is useful for visualizing the current wing design to verify that the geometry is acceptable on aesthetic and packaging standpoints. Buttons allow the user to browse various angles of attack and cycle through different Reynolds number airfoils with ease. The user is also able to select between graphs of CD vs. CL, CL vs. , CD vs. and CD/CL vs. depending on the desired lift and drag characteristics of an airfoil. Excel parameterizes the planform layouts and reduces the design problem to four variables with given geometric constraints: wing span, root chord, tip chord and sweep angle. (See Table 4) Since the initial optimization objective is to find the lowest possible drag for reasonable geometric constraints, overall wing weight is a secondary factor in comparison to drag optimization, and thus is not included as one of the constraints. Airfoil Selection: Using the Matlab GUI with the Selig data we can narrow down the 650 airfoils to 9. This is done by selecting airfoils that have high CL/CD values at the design lift coefficient of 0.55 and a Reynolds number between 40,000 and 65,000. As Reynolds number increases, the CL/CD values required for a feasible airfoil increase due to the need for a lower aspect ratio and the resulting increase in induced drag. This results in fewer airfoils that meet our goal as the Reynolds numbers increase. With secondary optimization of the nine airfoils selected, six have favorable characteristics. Only one of these surpasses the others in both low drag and low weight (see Table 5). Airfoil gm15 makes possible a planform of low weight, short wing span and low drag. Note that sweep value refers to the sweep angle added in addition to the sweep induced from the taper ratio. The actual leading edge sweep angle will be higher than this value. Based on the geometric properties for our selected airfoil, our calculations yield an aspect ratio of 8.22 and a wing platform area of 0.525 m2. The aspect ratio is defined as b2/S; for a given wing area, S, a large aspect ratio means a large span. From a drag Wing Drag(N) Mass(kg) b(m) Sweep croot (m) Ctip (m) gm15 s6063 s7012 rg14 rg14 gm15sm gm15sm sd7003 sd7003 2.4442 2.9398 2.5381 2.8761 2.8331 2.6638 2.4501 2.5914 2.5313 0.82 1.86 1.41 2.22 1.33 2.55 1.59 2.62 2.34 2.078 3.1 2.922 3.645 2.848 4 3.162 4 4 4.9 7.1 4.9 7.1 5 6.2 4.6 6.6 4.8 0.422 0.966 0.627 1.131 0.628 1.087 0.631 1.164 0.84 0.084 0.193 0.125 0.226 0.126 0.217 0.126 0.233 0.168 Table 5: Multiple Optimization Results for 6 wings 5 Mean Aerodynamic Chord & Center Mean aerodynamic chord ( c , M.A.C.) is a parameter, directly associated with the Reynolds number ( Vc / ). Equation ( 6 ) shows the M.A.C. lift at the tip. A sharp edge makes it more difficult for the air to flow around the tip, because the flow often separates at these edges. A winglet or endplate blocks the flow from the bottom to the top of the wing. This offers the greatest benefit to low aspect ratio wings whose wing tip vortex is strong, by increasing the effective wingspan, which further decreases induced drag. Our concern with using a winglet is the additional wetted area, which will create a larger parasite drag nullifying any reduction in induced drag. An unswept wing tip curves upward to increase the effective wing. We are using this design since it is similar to adding a winglet, without an increase in total wetted area. as a function of the taper ratio () and the exposed root chord (CR). M . A.C. 2 C R (1 ) 0.307 3 1 (6) Y, the distance of the M.A.C. from the centerline of the aircraft is 0.436 meters. This distance is dependent upon the taper ratio and the wing span. Aerodynamic center is the point on the aircraft where the airfoil pitching moment is constant with a change in angle of attack. It determines where to position the wing, and is important in stability calculations. In subsonic flow, the aerodynamic center is typically located at the quarter-chord point on the mean aerodynamic chord line, which is found to be 0.0768 meters from the leading edge of the wing. From equations historically used in aircraft design, such as Equation ( 7 ), we conclude that the aerodynamic center will be located 0.703 meters from the nose of the UAV. A.C. 1.5CR 0.25c Wing Weight and Structural Considerations: With any extraterrestrial mission weight is a top priority. Our mission is no different. Generally, wing weight selection is an iterative process between a structures team and an aerodynamics team. Due to time constraints, we approximate the wing weight with equation ( 8 ), a formula based on historical data used by many aircraft manufacturers as an initial wing weight. The formula is a relationship between, dynamic pressure, q, aspect ratio, AR, total aircraft weight, Wdg, thickness-to-chord ratio, t/c, load factor, n, taper ratio, , wing area, Sw, sweep angle, , and a multitude of constants, C1 through C14, that have been obtained using years of data for three classes of aircraft: fighter, transport and general aviation. Using the general aviation constants and the planned weight of the craft, 10 Kg, a circular reference of wing area, lift force, total craft weight and wing weight is created in Excel. Turning on the iteration command in Excel causes the values to converge to a steady state solution for weight analysis, solving with ease a process otherwise overly complicated by hand. (7) Wing Vertical Position The wing’s vertical placement with respect to the fuselage can be at three locations: atop, below or through the middle of the fuselage. A high wing design is used primarily for cargo planes, allowing the fuselage to be placed closer to the ground. However, the passing of the wing box over the fuselage will increase the parasite and pressure drag due to the increase in frontal area. This increase in frontal area is also present in a low wing design, which is used by virtually all commercial transport aircraft due to the advantages in landing stowage. Since we are not concerned about landing, we have no need for a low wing design. The advantage to a mid-wing design is that it gives the UAV more maneuverability while having a lower frontal area than the high or the low wing design. Due to the advantages associated with a mid-wing design, we are placing the wing in the middle of the fuselage. c C8 Wwing C1C2C3WdgC4 nC5 S wC6 AC7 t ...C9 C10 Wing Tip Selection A wing tip can prevent high pressure air beneath the wing to “escape” around the tip of the wing to the low pressure region above, resulting in a loss of lift at the tip of the wing, which is highly undesirable. Four different wing tip designs are considered: rounded, sharp, winglets and unswept. A smooth-rounded tip is precisely what we want to avoid. It easily permits air to flow around the tip and reduces (a) cos C11 S C12 f C13 q W (8) C14 fw (b) Figure 4: (a) Upswept Wingtip (b) Inverted V-Tail 6 them. Plotting the CL/CD vs. angle of attack () data for all four, we find the tail desired angle of attack to be between 3 and 5 degrees. By comparing the values of CL/CD at 3 degrees for each candidate, it is apparent that SD7003 is the best airfoil at the value we need for angle of attack. Tail Arrangement Selection Tails act as small wings; their purpose is to provide trim, stability and control to the craft. Trim is the generation of the proper lift force to balance pitching moment about the center of gravity. Stability and control are the tail’s ability to restore the aircraft from a perturbation in pitch, yaw, and roll, which is discussed in detail in the section of stability and control. There are a variety of possible aft-tail arrangements. Our design focus is on reducing parasite drag. We can thus narrow our search to four possible tail configurations based primarily on the wetted area of each tail configuration: conventional, V-tail, inverted V and Y-tail. Conventional tails, used on over 70% of all aircraft, have the typical vertical and horizontal tail seen on most commercial airliners. It provides adequate stability and control at a reasonably light weight. With the V-tail, as the name suggests, the vertical and horizontal tail components are combined in an attempt to reduce the wetted area. The horizontal and vertical forces, on the V-tail, are the resultant of their respective projections from the two angled surfaces. In order to provide the proper movement, the rudder and elevator on a V-tail are combined to create “ruddervators.” The problem with a V-tail is the production of a rolling moment in opposition to the desired direction of turn, known as “adverse roll-yaw coupling.” This produces a spiraling tendency when the UAV is making a turn. The inverted V-tail avoids this problem; it instead produces a desirable “proverse rollyaw coupling”. The “Y-tail” is similar to the V-tail, with a reduced dihedral angle and a third surface mounted vertically beneath the V, giving the UAV more yaw control. A drag penalty involved with adding another control surface causes the Y-tail design, like the conventional design, to not fit our design goals of minimizing drag. Using an inverted V-tail gives us the low drag required with greater stability than the standard V-tail. Stability and Control Stability and control is an integral part of designing an aircraft. It is vital that the aircraft is stable and able to handle moments, from various disturbances, while maintaining control. An aircraft possesses three degrees of freedom, pitch, roll, and yaw, and has two types of stability, static and dynamic. A system is statically stable if forces and moments acting on a body, as a result of a disturbance, initially act to return the body towards its equilibrium position. A system is dynamically stable if it eventually returns to and maintains its equilibrium position over a period of time. For our case, our top concern is longitudinal static stability, involving the pitching moments about the center of gravity. Though, as with any aircraft, lateral-directional static stability and control involving yaw/roll moments are also important. The steps to design a stable aircraft are as follows: 1) Make an assumption for the location of the center of gravity with respect to the nose of the aircraft 2) Make an educated guess regarding the placement/sizing of the tail 3) Determine the moment about the center of gravity due to the wing, fuselage, tail, and payload 4) If the moment coefficient, calculated in equation ( 9 ), at zero lift (Cm L=0) is positive and the slope of the moment coefficient versus angle of attack (dCM/d) is negative, the aircraft is longitudinally and statically stable 5) Reiterate the process if necessary 6) Determine the static margin Tail Airfoil Selection The goal in selecting an airfoil for the tail is similar to the wing, in that we want an airfoil with high lift-to-drag ratios. Since tails are small wings themselves, we look at the final low Reynolds number airfoils from the wing selection process. The main purpose of tails is not to generate lift, but to provide stability and control. For this reason, airfoils used for tails typically have little to no chamber to them. Since our wing airfoil has high camber, the chosen wing airfoil, gm15, is not applicable for the tail. Based on lift and drag data, as well as amount of camber, we limit our choices to four airfoils: SD 7003, S6063, S7012, and RG14, all of which have only slight camber to CM M cg 1 2 V 2 Sc (9) Equation ( 9 ) solves for the moment coefficient about the center of gravity, where M is the moment contribution about the center of gravity, is the density of the atmosphere, V is the cruise speed, S is the wing platform area and c is the mean aerodynamic chord, dependent on taper ratio. We use equation ( 10 ) to determine the moment coefficient about the center of gravity at zero lift; recall that this value must be positive for the aircraft to be stable. The first term on the right hand 7 side of the equation is the moment contribution from the wing and the fuselage, about the center of gravity. The second term is the moment contribution from the tail. CM(payload) is the moment induced by the payload. The moment coefficient about the center of gravity from the wing and fuselage, CM,CGwb, is approximated as the sum of the moment contribution of the wing body about the aerodynamic center and the moment generated by the lift force from the wing, M cgw M acw LW cos w (hc hacw ) . Drag terms are St, to be 0.15 square meters and the tail moment arm, Lt, to be 1 meter. Solving for tail moment coefficient yields a value of 0.25, which more than compensates for the moment coefficient from wing, fuselage and the payload. dCm d h hacwb Vh t / 1 d d With the first stability criteria satisfied, we can solve equation ( 12 ) for the second criteria, negative dCM/d Where and tare the lift curve slopes of the wing and the tail respectively, the quantity (h - hacwb) is the distance between the center of gravity and the aerodynamic center, and d/d is the rate of change of the downwash angle with respect to the angle of attack for the wing, approximately 0.45. dCM/d satisfies the second stability criteria with a value -0.03 With both of the criteria satisfied the UAV is longitudinally and statically stable. The neutral point is a fixed point on the UAV behind the center of gravity where dCM/d is equal to zero, and must be aft of the center of gravity in order to achieve longitudinal stability. Setting dCM/d to zero and solving for h gives the location of the neutral point at 1.06 meters from the nose of the craft. The static margin, the difference between the neutral point and the center of gravity, is 0.217m. This parameter is directly related to the stability of the UAV, the larger the static margin, the greater the pitching moment must be to cause a change in the pitching angle. However, too large a static margin may cause the flight controller to go unstable due to unacceptably high reaction latency. On the other hand, too low of a static margin will yield an aircraft that is inherently unstable in regards to pitching motion, and will require very fast control response to maintain steady state cruise conditions. Our static margin is between these two extremes, allowing stability without an over-active controller. In many ways, the lateral-directional analysis resembles the longitudinal analysis. Lateral-directional stability is the tendency of the UAV to return to a wing-level attitude after being displaced from a level attitude by roll or yaw moments, from such things as turbulent air. There are two primary factors for lateraldirectional stability: wing dihedral angle and wing sweepback angle. Dihedral angle is the angle at which the wings are slanted upwards from the root to the tip; not included in the calculation of wing and fuselage moment coefficient, as they are negligible compared to the other terms. Solving equation ( 10 ) gives us a value of 0.041 for CM,CGwb. C M ,cg L0 CM ,cgwb VHat (it , eo ) CM ( payload) ( 10 ) To determine the moment coefficient from the payload, we establish a series of point loads to approximate the weight distribution. This layout can be seen in Figure 5. According to our calculations the center of gravity will be 0.85 meters aft of the nose of the UAV. Moment calculations yield a moment coefficient value of -0.153 (negative moment being in the clockwise direction.) In order to satisfy the first stability criteria, a CM,cg greater than zero, the tail must be large enough to balance the clockwise moment produced by the payload and lift forces. According to our calculations the tail moment coefficient must be larger than 0.194. The moment coefficient is defined as Vhat(it+eo), where Vh is the tail volume coefficient, at is the lift curve slope of the tail airfoil (~ 0.1), it is the tail setting angle (~3 by standard convention), and eo is the downwash angle, which can be neglected. Vh lt St cw S w ( 11 ) Vh, the tail volume coefficient, is proportional to lt , the distance of the tail from the center of gravity, St ,the platform area of the tail, cw , the M.A.C. of the wing and Sw , the wing platform area. Equation ( 11 )shows that the further away the tail is from the center of gravity, the smaller the area of the tail needs to be. We want to make the tail large enough to give us adequate stability, but not as to further increase the parasite drag. To optimize both parameters, we set the tail platform area, 0.25 0.65 0.75 0.85 1.60 0m 4.2kg Battery 1.75 2.1 m cg 1.17kg Lidar Controls ( 12 ) 0.61kg Camera 1.8kg Motor Figure 5: Payload Point Mass Layout for Aerial Vehicle 8 0.6kg 0.43kg drive shaft Propeller its main purpose is to correct roll moments. The stabilizing effect of dihedral occurs when an aircraft sideslips slightly as one wing is forced down in turbulent air or during a turn. This sideslip results in a difference in the angle of attack between the higher and the lower wing. The increased angle of attack on the lower wing produces an increase in lift which helps the wing return to its level position. Research leads us to a dihedral angle of 3.5 degrees. Sweepback is the angle between the line formed from the front of the wing and the line perpendicular to the centerline, in the plain of the aircraft. The effect of sweepback in producing lateral stability is similar to that dihedral angle. A yaw moment increases the sweepback angle in one wing panel and decreases it for the other side of the aircraft. The change in sweep alters the effective dynamic pressure normal to the quarter-chord line of the wing panel, increasing the lift on one side of the wing, lowering it on the other side, and producing a restoring moment. Historical trends in wing sweep back give us a sweep angle of 5 degrees. Furthermore, being a 6 degree of freedom system, a soundly designed multi-input/multi-output, or MIMO controller to be used is required for an aircraft. The designing of such controllers are currently beyond our expertise. Cruise Performance The stall speed is determined directly by wing loading and the maximum lift coefficient. Stall speed is a major contributor to flying safety, indicating the minimum speed that will keep the UAV aloft. When an aircraft flies below the stall speed, flow around the airfoil begins to separate; as a result, a rapid loss in lift will be experienced. At that point, if the velocity is not increased beyond the stall speed, the aircraft will lose altitude and thus lose control. Equation ( 13 ) determines the stall speed where W is the aircraft weight, is the fluid density, S is the wing platform area and CLmax is the maximum lift coefficient of our airfoil, obtained from the graph of CL vs. angle of attack. Our aircraft has a stall speed of 83.4 m/s; we must fly above this speed in order to maintain adequate lift. Since our main goal is to map as much terrain as possible, it is necessary to maximize the range. To do so, we must fly at the speed where lift to drag ratio is greatest, given in equation ( 14 ) where CDP is the parasite drag coefficient and AR is the aspect ratio of the wing. Tail Geometry Having found the platform area of the tail, St based on stability constraints, it is important that we obtain the tail geometry. Determining the aspect ratio is crucial; having a large aspect ratio corresponds to a small chord, which further leads to an unusually low Reynolds number. Likewise, a small aspect ratio will lead to a substantial increase in induced drag. After much consideration we decided on an aspect ratio of 4. Using a taper ratio similar to that of the wing, 0.35, we determined the length of the tail root and tip chord. Similar to that of the wing, the tail mean aerodynamic chord ( c t ) is a function of the root chord and the taper Vstall V L / Dmax ratio. Based the c t value, we determined the tail Reynolds number, which is approximately 30,000. This value is in fact only half of the optimum Reynolds number suggested for our selected airfoil (SD7003). However, because the tail angle of attack will be small, boundary layer separation would almost be nonexistent. Therefore it is ok for the operating Reynolds to be smaller than the optimal value. 2W SClMAX 2W S C DP AR e ( 13 ) ( 14 ) Range is maximized when VL/D is 130 m/s (Mach number of 0.56). If the cruising speed is set to be greater than this value, then there will be a substantial increase in drag. Values deviating from this velocity will result in a loss of range. Turning Performance Maneuverability plays an important role in the design of an aircraft. Unlike combat planes that perform sharp turns, our UAV only needs to be able to slowly turn to avoid physical obstacles that it may encounter. A crucial parameter in trying to figure out the turning performance of the aircraft is the load factor, n, defined as the ratio of lift-to-weight. In our case, n is approximately equal to 1.095. The wing bank angle, defined as cos-1(1/n), is approximately 24 degrees. Ailerons control the wing bank angle by equal and opposite deflection of the two wings trailing edges, one up and one down; thus increasing lift on one side and decreases lift on the other side of the aircraft, Control The primary aerodynamic controls available are ailerons, elevators, and rudders. Because we are incorporating a V-tail design, the functions of elevators and rudders will be combined into one. Ailerons are the primary roll-control device, which operate by increasing lift on one wing and reducing it on another. They range from 50 to 90 percent of the wingspan and 20 percent of the wing chord length. Since the aircraft is unmanned, a suitable, redundant controller must be designed to stabilize and maneuver the aircraft. 9 CLvs. Angle of Attack & CL/CD vs. Angle of Attack inducing a roll moment about the centerline. A rolling moment banks the airplane and tilts the lift vector to one side. The horizontal component of the lift vector accelerates the aircraft laterally, thereby curving the flight path. Equation ( 15 ) solves for the minimum radius of turn, R. V g n 1 13.4 1.0 ( 15 ) 13.2 13.0 CL 12.8 0.6 12.6 12.4 0.4 Because of our relatively small load factor, n, we obtain a value of 10.5 km for our turn radius. The UAV is not capable of performing a sharp turns; this is not a problem as we will have adequate warning for any turns that need to be made and the controller can take into account minimum turn radius. Equation ( 16 ) describes the turning rate (degrees/time) for the UAV, how large your turn is, in degrees, per unit time. d g n^ 2 1 dt V 13.6 1.2 0.8 2 2 13.8 12.2 0.2 Cl SELIG Data 12.0 Cl/Cd 0.0 11.8 0 0.5 1 1.5 2 2.5 3 Angle of Attack 3.5 4 4.5 5 Figure 7: CL vs. for CFD (Green) and Selig (Red)data CL/CD vs. for CFD (Blue) ( 16 ) outlining the 2D airfoil obtained from Professor Selig’s database and resizing it for our chord length. Once FLUENT reads the mesh we specify the atmospheric and flying conditions. In this case the closest to Martian atmosphere that can be used is an environment of carbon dioxide. In FLUENT we set the fluid properties to the values from Table 1 in the section on atmospheric data and a gravitational constant of 3.72 m/s2, roughly 4/10ths the magnitude of gravity on earth. Boundary conditions are set that specify the pressure far from the airfoil and the velocity in terms of x-y components and Mach number. The xy components allow us to vary the angle of attack without the need of creating a new mesh for each angle of attack we want to test With fluid and environmental properties set we determine the appropriate method to use to perform calculations. Since we are flying in with a low Reynolds number we use a laminar boundary regime. To verify our assumption about the Selig data it is necessary to run simulations at several angles of attack. With a chart of this data we can compare the CFD data with that from Selig and find that, while they do not lie directly on top of each other, they both yield an acceptable coefficient of lift for an angle of attack between two and three. The CFD data gives slightly higher CL values than the Selig data. Since we are backing up an assumption based on experimental data, not determining values, this inconsistency is acceptable. If this error were on the side of lower lift we would have to do more calculations to make sure there is no problem. As this is not the case, it stands that our assumption regarding the validity of the Selig data is acceptable. Fluent Analysis (Selig Data Verification) It is stated in the aerodynamic assumptions section that Professor Selig’s data for his numerous airfoils are correct. Using Fluent, a computational fluid dynamics (CFD) package, we justify this assumption with appropriate calculations. While the Selig airfoils are normalized by Reynolds number, Martian conditions vary greatly from those in Professor Selig’s wind tunnel. The first step to CFD is creating a mesh containing the airfoil in a large space with boundary conditions to simulate the Martian environment for our airfoil. This is done, using GAMBIT, from grid points Figure 7: FLUENT Pressure Gradient Around Airfoil (Red = High Pressure Blue = Low Pressure) Figure 6 10 C L /C D R 1.4 . Propeller Design Overview From past Mars aircraft concepts and high altitude, low speed Earth aircraft, propellers have been the preferred choice. Our choices for powering an electric motor are: batteries, fuel cells and solar cells (see discussion on solar cells in the Electrical section.) As a result of the lower speed of sound on Mars, due to low temperatures, and density about one-hundredth that of Earth’s, our effort focuses on generating the necessary amount of thrust, as well as keeping the tip speed of the propeller below supersonic conditions. If the tip Mach number reaches 0.85, there will be a large drop in the propeller efficiency due to the flow separation and formation of shockwaves. Propulsion Design Introduction Low atmospheric density and the lack of appreciable amounts of atmospheric oxygen complicate the propulsion for a Mars airplane. These constraints lead to the consideration of propulsion options that are more restrictive than those of Earth. The analysis carried out in the propulsion section of this report is based on an airplane that is not landing intact on the surface of Mars once flight is completed; if an airplane is intended to land or take-off from Martian soil, a new set of design specifications need to be considered. Propulsion Selection Since the use of a combustion engine is not feasible due to the lack of oxygen, our choices for the propulsion subsystem are limited. There are two methods to propulsion we consider for Martian aerial flight: chemical propulsion and propeller driven propulsion by an energy source. The use of monopropellant rocket thrusters enables combustion without the need for atmospheric oxygen, by carrying chemical compounds that burn spontaneously when ignited. It provides the UAV with uniform thrust; however, once ignited, the process cannot be stopped until the fuel runs out. Bipropellant thrusters, on the other hand, carry fuel and oxidizer separately. They are more practical in this case since the thruster can be turned on or off in order to maintain cruise speed at V(L/dmax). Bipropellant thrusters, however, tend to be more complicated to design. The thrusters found for our design constrains are capable of generating anywhere from 5 to 20 N of thrust and have a specific impulse, Isp, in the range of 300 to 350 seconds. Isp, a key performance parameter for rockets, is defined as the thrust that can be obtained with a propellant weight flow of 1 unit per second. Modern large scale rockets, like the one found on the Shuttle, can achieve a maximum Isp of around 450 s. The second approach to propulsion we consider is the use of energy from an on-board battery, nuclear device or an off-board energy source, such as solar energy, to power a propeller. Solar powered airplanes must have a large projected area to collect sufficient solar power and are inefficient when the solar intensity is low, as it is on Mars. Solid rocket propulsion is inherently simple; yet, as mentioned before, there is no way to control the thrust once ignited. Bipropellant thrusters run the risk of explosion due to low atmospheric temperature on Mars. Our calculations also indicate that for the same weight, a battery driven propeller would yield a much greater range than for a bipropellant thruster, so we select propeller as our form of propulsion Propeller Placement Examining the advantages and disadvantages of propeller placement along the fuselage places the propeller at the rear of the fuselage. The main advantage in using a pusher is in the aircraft’s capability to fly in undisturbed air. With a tractor propeller, the aircraft flies in the turbulence from the propeller wake, which could lead to additional drag. Drag Calculations The UAV will be operating at steady, level flight, where all the forces will be in equilibrium, meaning that thrust must balance the drag in order to keep the UAV at a constant cruising speed. C D C DP C DI C DC ( 17 ) Equation ( 17 ) calculates the total drag on the aircraft, where CDp is the parasite drag coefficient (also known as skin friction drag coefficient,) CDi is the induced drag coefficient and CDc is the drag due to compressibility. CDc becomes significant when the craft approaches sonic condition. However, because the UAV will operates at subsonic speed, CDc is be neglected. C DP ,wing C f k swet ( 18 ) 4 The parasite drag, coefficient from the wing is solved in equation ( 18 ), where Cf and k are roughness constants based on the wing Reynolds number and Swet is the total wetted area for the wing. Similarly, we are able to determine the parasite drag coefficient for the tail and the fuselage. Assuming that the wing, fuselage and tail contribute to 95% of the total skin friction drag, the overall parasite drag coefficient is determined to be 0.02225. 11 addition, the use of propfans, which feature 8 to 10 wide, short blades of sweptback planform are considered for blade configurations. If a propfan blade configuration can be utilized, being powered by an electric engine, opposed to the standard turboprop engine, it would be an option worth considering. Although counter-rotating blades and propfans were investigated as ways to improve propulsion, the aforementioned roll moment is beyond our current level of expertise. 2 C DI C L 0.016 Ae ( 19 ) Using equation ( 19 ) we solve for the induced drag coefficient, where CL is the wing lift coefficient, A the wing aspect ratio, and e the Oswalt efficiency factor. Combining the parasite and induced drag coefficients gives an overall drag coefficient of 0.0394. Thus we can determine total drag force experienced by the aircraft, from equation ( 20 ), to be 3.14 N. 1 D C D V 2 S 3.14 N 2 D prop 0.54 Pengine 2.27m ( 20 ) Propeller Efficiency As noted earlier, in the section it is essential that we keep the tip speed of the propeller under M = 0.85 or 195 m/s. The helical speed, the tip velocity on a moving aircraft, is the sum of the rotating speed at the tip of the propeller and the freestream velocity, calculated in Equation ( 23 ) Motor Power Specification With the thrust required for level flight known, we specify the amount of power the motor needs to produce in order for the propeller to generate that much thrust. TVo prop Pengine ( 22 ) V ( 21 ) We determine the engine brake horsepower, Pengine, where T is the thrust required to maintain level flight, V is the flight velocity and η is the efficiency of the propeller. Since the UAV will be cruising at V(L/Dmax) the flight velocity is 130 m/s (see aerodynamic section for more details.) For a propeller efficiency of 95%, The engine will produce a power of 430 W or 0.58 HP. Although calm flight conditions are assumed, it is highly likely that there will be significant wind gust that will increase the drag value. Taking this factor into account, we impose a safety factor of 1.2. The propeller must therefore be capable of generating 3.77 N of thrust if necessary to maintain leveled flight; this corresponds to a maximum engine power of 490 W or 0.675 HP. Vtip V freestream ( 23 ) Vtip nD ( 24 ) 2 tip helical 2 The stationary tip velocity is calculated in Equation ( 24 ),where n is the rotation speed in revolutions per second and D is the diameter of the propeller. With a helical tip speed of 195 m/s, we obtain propeller rotation speed of 20 revolutions per second, or 1200RPM. The overall propeller efficiency is expressed in terms of the advance ratio v/nD, thrust coefficient CT, and power coefficient CP in equation ( 25 )), where CT and CP are defined in equations ( 26 ) and ( 27 ). The resulting propeller efficiency of approximately 93% is very close to the 95% value we assumed initially in calculating power. Propeller Diameter Using Equation ( 22 ), we determine the diameter of the propeller from the brake horsepower of the engine. Note, this diameter is equivalent to the length of our wing span; as the propeller rotates an induced roll moment is generated. We further explore the possibility of using counter-rotating blades. Counter-rotating blades have mainly two advantages: they are more efficient at high Mach numbers than a single propeller configuration and they allow a smaller diameter blade, allowing them to spin at higher RPM without a loss in aerodynamic efficiency. By having counter-rotating blades, our propeller diameter will be decreased considerably while maintaining the same efficiency. However, the extra blades will increase the weight compared to the use of a single propeller. In CT CP 12 v CT n * D CP T 0.0188 n 2 D 4 P 0.056 n3 D 5 ( 25 ) ( 26 ) ( 27 ) Propeller Pitch The pitch is the theoretical distance the propeller will advance along the axis of rotation in one complete revolution. There are two types of propellers: fixed pitch and variable (controllable) pitch. In a fixed-pitch propeller, the pitch is set by the manufacturer and cannot be changed by the pilot. There are two types of fixed pitch propellers: the climb propeller and the cruise propeller. The climb propeller has a lower pitch, which therefore leads to less drag. This results in the capability of higher RPM and more horsepower being developed by the engine; such will increase performance during takeoffs and climbs but decrease performance during cruising flight. On the other hand, the cruise propeller has a higher pitch and therefore more drag which results in lower RPM and less horsepower capability. Performance during takeoff and climb is therefore decreased; yet, efficiency during cruising flight is increased. Contrary to the fixed pitch, a variable pitch propeller permits the pilot to select a pitch that will result in the most efficient performance for a particular flight condition. Since we are solely dealing with cruise flight conditions, we select the fixed pitch cruise propeller for its simplicity and performance. blade will be traveling at different speeds. A small twist in the propeller blade must be incorporated to ensure that each section advances forward at the same rate which stops the propeller from bending. Designing propeller blades takes a great amount of expertise and years of experience. Since it is beyond our level of expertise, we will not determine the exact pitch, shape, twist, and airfoils for the propeller. Propulsion Summary As previously mentioned, the power consumed by the propulsion system will be 430 W/hr. The 4.2 kg of battery that is carried onboard will generate a total of 1050 W, which allows the UAV to stay aloft for approximately 2.3 hours. At a cruising speed of 130 m/s, this corresponds to a range of 1076 km. Structural Design Introduction In structural aspects, the objective is to design and verify the safety, stability, and reliability of the unmanned aerial vehicle. Both the wing and fuselage will be hollow in order to minimize weight and house instrumentation needed for the mission. The structure must also be able to withstand outside forces as well as its own weight. Propeller Blade Design In propeller design, deciding the number of blades to incorporate is essential. An optimization among efficiency, thrust and weight shows that a threebladed propeller is preferred. Not only is it capable of producing more thrust than would a two-bladed propeller, it is also lighter and more efficient than using a four-blade propeller. The tip section of the propeller revolves faster than the root section; therefore, the Reynolds number along the propeller changes as the radius increases. As a result, one would have to select a different airfoil for each section of the propeller blade. Structural Design/Fabrication Two procedures are considered in designing the structural body of the unmanned aerial vehicle. The first process consists of making a skeleton with trusses and placing coatings of outer layers on top of this system to form the external shell. The second option is to make a mold out of Styrofoam or a similar solid foam material in the exact shape of the aircraft and then coating the outer layers of skin on top of the mold. After the layers are set, the inner mold is removed and only the thin shell remains, but shaped in the form of the aircraft. The option implemented in this project is the latter process of coating a mold. This is primarily due to weight considerations. Although trusses increase the weight of the aircraft, a few trusses will be used for support. The trusses will serve solely for structural purposes, and not to form the shape as do the trusses in the first process. The coating of the aircraft will consist of three layers: a base layer, a middle body layer, and a surface finish. The materials chosen for the first two layers must have a low density to optimize weight, but must also be structurally sound. In addition, the materials must be capable of withstanding the extreme temperatures in the Martian environment, which reach on average -63°C on the surface. The material picked Figure 8: Propeller Blade Illustration The optimum airfoil thickness will be around 15 to 18 percent near the root, progressively thinning to 10 percent at the tip. Propeller blades are in fact wings themselves, producing a resultant aerodynamic force that may be resolved into a force pointing along the axis of the airplane. Thus, similar to the airfoil for our wings, the blades should have a high aspect ratio in order to minimize drag. An elliptical-based shape blade with a rounded tip would yield optimal performance. As the propeller spins, each section of the 13 for the bottom layer is 0.1mm aluminum. The body layer will be 0.5mm carbon fiber, and the outer layer will be polished aluminum. Wing Several approximations are made in designing the wing of the aircraft. First, the wing is treated as a cantilever beam and second, as being hollow. The following sections will explain further analyses on the wing. Figure 9: Wing Model The thickness at the center of gravity, “y”, is also highly approximated—the front face of the wing is modeled as a 2-dimensional trapezoidal figure. From our trapezoidal face, we can find a linear relationship of thickness, y, to position along the length of wing, x. Wing: Material Considerations In choosing the wing material, the following parameters are of importance: Low density, ranging from approximately 1300 kg/m3 to 1700 kg/m3 High fracture toughness (FT) approximately 1.0 x 105 Pa-m1/2 High tensile strength that is capable of: a) Supporting instrumentation weight b) Resisting forces of lift, 44.0 N c) Resisting forces of drag, 3.20 N d) Resisting forces of gravity Usage temperature between -80oC and 40oC Easily molded, shaped, and machined y@root y@tip x y@root 0.0116 x 0.0185 y l total wing Upon determining the horizontal component of the center of gravity, we can find the thickness at this point, y. Values can be found in table 7. Carbon Fiber Aluminum Center of Gravity (m) Thickness (m) Base (root/tip) (m) Height (root/tip) (m) Length (m) Y (m) 1500 0.4342 0.0005 0.422 0.037 1.039 0.0135 2700 0.4342 0.0001 0.422 0.037 1.039 0.0135 I (m4) Total Volume (m3) 1.781e -6 2.469e-4 1.781e-6 2.469e-4 Mass (kg) Force, weight (N) Force, lift 0.3704 1.370 22.0 0.6667 2.467 22.0 Moment (Nm) Stress (Pa) 8.957 8.347e5 8.481 7.903e5 Material Density(kg/m3) Given these material parameters, we identify carbon fiber to be optimal; its material characteristics are shown in Table 6: Property Value Density (kg/m3) 1500 Fracture Toughness (Pa-m1/2) 5.7 Tensile Strength (MPa) 13.9 Young’s Modulus (GPa) 71 Hardness-Vickers (HV) 42 Temperature, min (C) Temperature, max (C) ( 28 ) -273 2002 0.152 0.0129 0.152 0.0129 Table 7: COG thickness and Aluminum Stress Analysis Table 6: Carbon Fiber Properties Wing: Stress Analysis It is crucial that the structure of the UAV be strong enough to withstand aerodynamic forces while in flight. We therefore perform stress analysis on the wing to verify its capability of flying in Mars. The overall wing dimensions are known based on calculations from aerodynamics; the thickness of the hollow wing is dependent upon the yielding of carbon fiber. The wing is approximated as a cantilever beam with its airfoil modeled as a rectangular cross-section, as seen in Figure 9. Varying wing thickness corresponds to varying values of mass—and therefore, varying values of gravitational force. Because the aircraft is lightweight, the force of gravity is nearly negligible. This force of gravity and 44 N force due to lift provide the moment used to determine bending stress. The moment of inertia for the approximated rectangular cross-section is determined at the center of gravity. I I outer I inner 1 3 bh 3 b 2t h 2t 12 ( 29 ) For a wing of 0.5mm thickness, we can calculate the bending stress. 14 My 0.8347 MPa I ( 30 ) Thus, since the yield stress of carbon fiber is known to be 300MPa, we can conclude that a wing of 0.5mm thickness will not yield. Performing similar calculations for the inner aluminum core layer, we calculate a bending stress of 0.7903MPa. In comparison to its yield strength of approximately 20MPa, we confirm that the aluminum will not yield. Results are depicted on table 7. the fuselage; the batteries would therefore be on the top and mid floor. Fuselage: Thermal Design (Note: all the equipments will be placed inside the fuselage; therefore we are solely concerned about designing a thermal control system for the fuselage only) The main purpose of having a thermal control system is to maintain batteries and instruments at their optimal operating temperature. There are two basic approaches to the design of a spacecraft’s thermal control system – passive and active. Passive control operates by using appropriate materials and surface finishes so that the fuselage temperature remains within acceptable range of temperatures. The latter uses mechanical or thermoelectric devices. For instance, the UAV would consist essentially of a central thermal transfer bus, a fluid loop transporting the heat from the radiator to the individual components. We choose a passive thermal control system because it is more reliable and easier to design than an active one. The ideal operating temperature range for our instruments is approximately 270-290 K. Due to the extremely low temperature on Mars and the fairly high velocity that we will be flying at, heat losses due to conduction and convection would be significant. Therefore, we want to incorporate a high-absorptance and low-emitting metallic surface. This would allow the UAV to trap as much heat from radiation as possible while emitting relatively a little amount. The optimal fuselage equilibrium temperature would be obtained by varying values of and through the following heat balance equation: Wing: Buckling After determining that the wing will not yield, it is important to confirm that the wing will not buckle. Pcr, critical load, is found at the tip of the wing, where buckling is most likely. The load that the UAV will experience due to lift and gravitational force of each material is approximately 26 N, well under the critical load of 1613 kN. Thus, the wing will not buckle. Pcr 2 EI L2 1.613e6 N ( 31 ) Fuselage Once we have established the dimensions of the wing, we must now consider the design of the fuselage. The shape of the fuselage will be one similar to sailplane design, for weight minimization and aerodynamic purposes. Fuselage Interior In order to accommodate the various instruments required in this mission, the interior of the fuselage will consist of three floors. These floors will be made of honeycomb sandwich structure (Figure 10). Among the advantages of using honeycomb are lightweight, high crush strength and stiffness, structural integrity and high fatigue resistance. As / cTeq Qconduction Qconvection 4 AsJ s APJ a APF12 J P Qint ernal ( 32) The left hand side of the equation consists of the heat that is escaping the system. The term As/cTeq4 is the heat loss due to radiation. Q conduction and Qconvection are heat losses due to conduction and convection, respectively. As/c is the total surface area of the UAV. is the Stefan-Boltzman constant (5.67x10-8 W/m2K4). The right hand side of the equation is the sum of the heat being absorbed by the system. The term AsJs is the heat addition from solar flux of the sun. ApJa is the heat addition due to planetary albedo. ApF12Jp is the heat addition from planetary radiation. Finally Qinternal is the internal heat generated by the various instruments. {As and Ap are the projected areas of the UAV in the directions of the sun and the planet respectively. Js is the solar radiation intensity on Mars (590 W/m2). Ja is the albedo contribution to the total radiation input to the UAV, which is defined as the product of Js, a (average Mars Figure 10: Honeycomb Structure The honeycomb structure material will be made of aluminum, due to its smooth and thin cell walls, in addition to the high strength-to-weight and stiffness-to-weight ratio. Particular instruments must be laid out in specific locations within the fuselage; for example, the LIDAR will be placed on the base floor. For the LIDAR to serve its purpose, this part of the fuselage base will have glass material as a “window.” The shaft and motor will be further back, towards the tail end of 15 albedo, 0.15), and visibility factor F, which is dependent on the altitude of flight and the bearing angle.} We can model heat loss due to conduction as a 1 -dimensional conduction problem. of the materials that we considered along with their properties: Figure 11: Model of heat loss as conduction The total heat transfer via conduction can be expressed in equation ( 33). Tatm Teq qx 1 L1 L2 [( ) ( )( )] hA KalA KcfA Nu L K L a/e Equilibrium Temperature (K) Polished Aluminum Surface 0.35 0.004 8.75 265 Polished Stainless Steel 0.50 0.130 3.85 n/a Polished Copper 0.28 0.130 2.20 n/a Grafoil 0.66 0.340 1.90 n/a Vapor-blasted Stainless Steel 0.60 0.330 1.80 n/a Gold/Kapton/Aluminum 0.53 0.420 1.260 n/a Gold-plate on Aluminum 0.30 0.040 7.50 261 Fuselage: Material Selection The fuselage itself is modeled as a pressure vessel. The entire structure is assumed to be sealed off when it is built; therefore, it would contain Earth’s atmospheric pressure (100 kPa). However, the ambient pressure on Mars is so low (700 Pa) that there would be significant pressure drop between the pressure inside the fuselage respect to the atmospheric pressure. Because air in a high pressure region tends to move to that of a lower region, there would be a tremendous force expanding outwards. The function of a pressure vessel is to essentially contain pressure P. The main objective is to do so while minimizing weight. A constraint involving this would be that the pressure vessel must leak before it breaks. This ensures that if a crack exists, the leak would release pressure gradually and thus safely. Based on these objectives and constraints, it is essential that we determine the appropriate material indices, which would help us assess the optimal materials. We can idealize the pressure vessel as a thinwalled cylinder with an average radius R and thickness t. The wall thickness is chosen so that at a certain pressure difference, the stress is less than the yield strength of y of the material. The stress should also be less than the fracture stress, at which point a crack would propagate in the vessel. It is however, notable that there would be no worry of crack propagation if the stress is kept under the yield stress—this ensures stable deformation. Such is expressed in equations ( 28 ) - ( 37 ). ( 34 ) ( 35 ) The Reynolds number is approximated to be 280,000. The Prandtl number is somewhere around 0.76. As expressed in equation ( 36), h is simply a function of the Nusselt number, conduction coefficient k of the ambient, and L the length of the flat plate. h e The temperatures of Steel, Copper, Grafoil, and Gold/Kapton/Aluminum of Table 8 are found to be well under 265K and therefore not applicable. Thus, based on the values and given equilibrium temperatures, in Table 8, our chosen external thermal layer will be one of polished Aluminum surface. Equation ( 34) expresses h as the convective heat transfer coefficient, A as the surface area of the fuselage, Tatm, Teq as the atmospheric and equilibrium fuselage temperature, respectively. In order to obtain the convective heat transfer coefficient h, we use the Nusselt correlation of equation ( 35) for flow over a flat plate. NuL 0.664 Re1/ 2 Pr1/ 3 a Table 8: Metallic Surface Materials - Property Table ( 33 ) where Tatm is the average ambient temperature (230 K) at our cruise altitude, h is the convective heat transfer coefficient, A is the surface area of entire fuselage, L 1 is the thickness of the aluminum (0.1mm), L2 is the thickness of the layer of carbon fiber (0.5mm), and K is the thermal conductivity of aluminum and carbon fiber respectively. After plugging in values for all the variables, the equation is reduced down to 1.4(230-Teq). Next, we model heat losses due to convection by the fuselage as fluid flowing pass a flat plate. The total heat transfer via convection is defined in equation ( 34). Q hATatm Teq MATERIAL ( 36 ) Now that all the terms in equation ( 32) are in terms of and Teq and Teq, we can select metallic surface finishes with reasonably and low values that would give us comfortable temperature for the instruments to be operating at. The following is the list 16 K ac C 1C y 2 M4 is expressed in Figure 13 —strong, light materials lie near the top of the figure. 2 ( 37 ) Maximizing the material index, M1, will maximize tolerable crack size. M1 K1C y ( 38 ) Leaks in the pressure vessel caused by a crack can be detected if the crack is just the size to penetrate the inner and outer surface while maintaining vessel stability. We must note that the wall thickness is such that it will not yield. Thus, a maximum value of the material index M2 indicates safe containment of maximum pressure. y M4 Figure 13: Specific Strength Chart of Materials ( 39 ) Based on these material indices, carbon fiber is chosen for the fuselage structure. While maximizing M1 and M2, we keep in mind that minimizing wall thickness is an important objective. Since small values of wall thickness indicate high numbers of yield stress, we try to maximize yield stress for our material. Thus, another material selection criterion would be to maximize index M3. M3 y Propeller Shaft A cylindrical shaft in the tail end of the fuselage interior will connect the propeller and the motor. Conventionally, the shaft is 1/20 of the fuselage diameter. We will perform calculations of torsion to determine the minimum size that is required of a shaft of carbon fiber. The total torque created by the propeller is approximately 5330 Nm; this is calculated by the force generated by the propeller, taking all three blades into account. Using the Tresca yield criterion, where max=y/2, we use equation ( 42) for maximum shear stress to find an optimum shaft radius. ( 40 ) Because our vessel is in fact an aircraft, weight is crucial. For minimizing weight, we must further find a material in which the material index M4 is maximized. M1 K1C y ( 41 ) max Once the four material indices have been found, a material selection chart is referred to. Figure 12 is one of fracture toughness (K1c) versus elastic limit (y). The diagonal line corresponds to M2, expressing the constraint that the vessel must leak before it breaks. T c3 2 ( 42 ) With a safety factor of 2, we result in a shaft diameter of approximately 0.0712m. Performing weight calculations based on density, we find that using the given diameter, the shaft will weigh 6kg—this exceeds our initial weight constraint. We further continue our shaft analysis by making the shaft hollow. C2 C1 Figure 14: Shaft Diagram Figure 12: Fracture Toughness vs. Elastic Limit 17 The equation for maximum shear stress is now illustrated in equations ( 43)-( 44). max J c 2 Tc2 J 4 2 c1 finish, the total mass is considered negligible. Weight is therefore calculated as the product of the surface area, the thickness, and the density of the material. A small area of glass fiber, to accommodate the LIDAR instrument at the base of the fuselage, is noted. Including the shaft and propeller as part of the structure, the UAV will have a structure of 2.97kg. Note Table 9 for specific weights. ( 43 ) 4 ( 44 ) Table 9: Specific Weights of Chosen Materials Performing calculations, we achieve a shaft radius of approximately 0.03m. For a safety factor of 2, this will increase to approximately 0.0815m. The mass is still too large, at around 1.73 kg. It is therefore necessary to decrease the length of the shaft to around 0.25m, in which case the mass will decrease considerably, to around 0.43kg. Thus, the motor must be designed to fit into a tube with a diameter of approximately 0.09m. Structures Conclusion For our given structure, bending moment and buckling analysis is done on the wing, in addition to pressure vessel analysis on the fuselage. We can therefore verify that our structure is both stable and sound—the wing will not yield nor buckle and the fuselage will not break given our dimensions. Our total structure weight achieved is approximately 3 kg. Surface Total Thickness Density Electrical and Systems Design Area Mass 3 (m) (m2) (kg/m ) (kg) Introduction: The0.0005electrical systems sub-team 0.827 1500 and 0.6203 concentrates on mapping the topography 1.200 variable 1500 0.6000 of Mars given weight parameters set by the structural sub-team. An 0.125 0.0005 1500 0.0938 electrical system needs to power and control the plane, n/a n/a 1500 0.4300data and flight as well as gather topographical n/a n/a 1500 0.6000 information. Listing flight-critical sensors and datagathering equipment is the first step in the design process. We compile a chart of mass and power consumption based on trade studies, then 0.827 0.0001 2700initial 0.2233 consider instrument networks to provide data storage 1.200 variable 2700 0.3600 and distribution capabilities. Once a general idea of 0.125 0.0001 2700 forms, 0.0338 equipment and power supply we reduce mass in light of flight time optimization by using lighter and more efficientn/asensors along with more 0.827 energy 0.0001 Negligible advanced energy sources. 1.200 0.0001 n/a Negligible The overall mass for the instruments is 7.80 0.125 0.0001 n/a Negligible kg with an energy consumption of 459 watt-hours. Fuselage Wing Tail Shaft Propeller Fuselage Wing Tail Fuselage Wing Tail Assumptions: 0.009 0.0151are made in the The0.0007 following2500 assumptions design of the UAV electrical systems: Fuselage 1. Overall Structure Weight Keeping the structure stable and lightweight is crucial for the UAV. Overall structural weight is based on the surface area and material thicknesses. The UAV can be divided into three sections, being the tail, wing, and fuselage. Each of these three sections will have three layers of material—the external thermal layer and inner aluminum core layer each being 0.1mm thick, and the structural carbon fiber layer being 0.5mm thick. Because the polished aluminum is merely a surface 2. 3. 4. 18 0.1000 TOTAL 2.9761 operating temperature inside The the UAV is greater than 0C. There are no directional magnetic fields present to adversely affect navigation. There will be a network of Mars-orbiting satellites as outlined by NASA. This will enable direct line-of-sight communication with the UAV at all times. All instruments are customizable for the UAV and Martian environment. Electronics will be radiation-hardened to ensure adequate 5. 6. performance during transit and on Mars. All wiring will be shielded to prevent electromagnetic interference and to reduce transmission losses. The mean-time-to-failure (MTF) of all instruments is much greater than the expected duration of the mission, which includes transit time and time spent on Mars. Manufacturer & Product Name Aeroflex RadHard UT80CRH196KDS Aeroflex ACT5108 RadHard Motor Driver Cognex MVS-8100D Digital Frame Grabber DSP Arch. DSP24 24-bit HP Digital Signal Processor Aeroflex UT28F256 LV PROM SEAKR NV-CPCI NonVolatile, Solid-State FLASH Constraints: The following constraints are imposed on the design of the UAV based on current technologies and design methodologies: 1. 2. 3. The total mass of all instruments must be less than 8 kg to accommodate the aerodynamic and structural constraints listed earlier. The data rate used for communications is limited to 480 MB/hr based on a two-antenna design and UHF frequencies. The oxygen-deficient atmosphere prevents the efficient use of fuel-cells as a potential power source Function Power MCU 0.48 W Motor 0.2 W Video 3W Comm 2.48 W Memory 1.5 W Memory 3W Table 11: Micro-controller Instrumentation Microcontroller Unit and Memory: Table 11 lists the components that comprise the main micro-controller unit. All circuit components are radiation-hardened to at least 300 Krads during the fabrication, design, and layout processes to ensure they perform as expected after a three to six month transit period and during the two hour mission on Mars. Electronic circuits are exposed to approximately 1.75 Krads[Si]/year in space when shielded with 50 mils of aluminum, with a slightly higher number (10 krads[Si]) during solar storms. Therefore, the radiation-hardened components are capable of withstanding several years of exposure without side-effects. However, radiation-hardening requires larger layout footprints and more conservative transistor designs, resulting in slower switching speeds and larger Instrument Selection: Table 10 lists the UAV instruments and their respective mass, power consumption, operating temperatures, and physical dimensions. The following sections discuss the trade-offs and specifications for each instrument. Important sections such as power management, communication, and navigation are considered in more detail followed by a high-level block diagram of the system. Name Mass (kg) 0.003 0.004 0.002 0.5 Power (W) 0.045 0.125 0.053 0.1 Op. Temp Dimensions (C) -40 to 125 7 mm x 7 mm x 3 mm -54 to 120 1.65 mm x 1.2 mm x 0.4 mm 0 to 85 18.9 mm x 17.5 mm x 7.6 mm > -5 variable Analog Devices ADT7317 (x3) Endevco 32394 Si MEMs Pressure Sensors (x2) Motorola MPXV5004G6U Low-Pressure Sensor Safety, navigation, control transducers (mechanical flaps, rudders, elevators) Analog Devices ADXL150 Accelerometer (x3) Analog Devices ADXL105 Accelerometer (x3) Analog Devices ADXRS150/300 Gyroscope Cognex CDC-100 CMOS CCD General Atomics LIDAR Receiver Triple-Junction Photovoltaic Compass Li-Ion SPE Batteries Microcontroller Instrumentation DC Electric Engine Patch Antenna (x3) Total 0.015 0.006 0.062 0.11 0.12 0.015 4.2 0.21 1.8 0.75 7.80 0.054 0.041 0.126 2.5 1 N/A -1050/hr 10.2 430 15 459/hr -40 to 85 7 mm x 7 mm x 3 mm -40 to 85 7 mm x 7 mm x 3 mm -40 to 85 7 mm x 7 mm x 3 mm 0 to 45 34 mm x 31 mm x 47 mm > -10 53 mm x 64 mm x 33 mm > - 10 90 mm (diam) x 240 mm > -20 140 mm x 140 mm x 140 mm 0 to 50 136 mm x 93 mm N/A > -10 50 mm x 50 mm 0C (min) Table 10: Instrument Selection 19 areas compared to commercial designs. The trade-offs are improved durability in harsh-environments and lower power consumption due to lower transistor densities - both of which are critical factors in the UAV design. The UT80 MCU is a 16-bit microcontroller designed to run on 20 MHz clock and industrystandard MCS-96 RTR architecture, which ensures compatibility of out-sourced software design. The 1 KB of internal SRAM is insufficient for all sensor and image data; therefore, the UT28F256 PROM and SEAKR FLASH external memory are added to augment data-storage capabilities. The UT28F256 external LV PROM adds an additional 256 KB of nonvolatile memory, which is used to primarily store the temperature, pressure, acceleration, and gyroscope sensor data using a sample rate of once per second. The SEAKR NV-CPCI FLASH chip allows for 1 GB of non-volatile memory with a maximum data rate of 27 Mbps and 3 W of power dissipation. The FLASH chip will be primarily used to store high-resolution images and LIDAR data. Table 12 lists the properties and memory bits required for each sensor and camera type. Extra bits are required for the Endevo pressure sensors to provide the required 8.62 mV/psi sensitivity. Similarly, due to the aerodynamic precision required once in flight, at least 10-bits of resolution is needed Sensor Type Quantity Total Bits ADXL150/105 3/3 60 bits @ 10-bits each ADXRS150/300 3/3 60 bits @ 10-bits each Endevo 32394 2 24-bits @ 12-bits each Motorola MPXV5004G6U ADT7317 1 8-bits @ 8-bits each 2 20 bits @ 10-bits each ThermalTab RTD 1 10-bits @ 10-bits each Total 18 182 bits (22 bytes) Software programs, such as navigation algorithms and communications protocol will be stored on the internal SRAM to allow for faster data-transfer rates to the MCU registers. LIDAR and Camera: The LIDAR unit, provided by General Atomics, includes a frequency-doubled Nd:YAG pulse laser, dual microchannel plate CCD detector, scanning mirror, and light-weight collection lens required for numerous high-resolution images. The unit is located in the nose of the UAV. In order to provide maximum coverage of the landscape, the scanning mirror, powered by a low-power DC motor, provides a fast horizontal scanning motion at several thousand points per scan. Based on the method developed by Lathrop et al., the CCD detector gathers all of the reflected laser intensity per scan. A horizontal resolution of much better than two meters is possible with this unit, which is a significant improvement over the LIDAR instrumentation onboard current Mars-orbiting satellites. The receiver specified by General Atomics provides a resolution of 320x320 pixels at 4 bits/pixel and a total of 50 kB/scan. The Cognex CDC-100 HiRes CCD provides 1280x1024 resolution at 8 bits/pixel of color with no compression, which requires a total of 1.25 MB/photo. The CCD interfaces with the Cognex MVS-8100D listed in Table 11. Power Management: Initial energy source selections for the UAV consist of (1) solar cells, (2) alternative electromagnetic sources, (3) fuel cells, and/or (4) batteries. The primary requirement for these energy sources is to provide adequate energy to on-board instruments and motor, specified by a minimum of 460 W for one hour. Moreover, they must function within Martian atmosphere; that is, they must function despite various gas compositions, temperatures, distances from the sun, and other factors. They must also be optimized in light of the ratio kilowatt-hour (energy) per kilogram. Trade-off analysis for the various selections shows that, first, solar cells do not provide enough required energy for flight alone or regeneration. This fact is expressed by a specific area of 263 W for 1 m 2 of a solar panel with a Martian efficiency of 28%, yielding only 73 W/m2. Note that given a wingspan of around 2 m x 0.5 m, only 73 W can be achieved from a full solar cell array. Additionally, since the energy per mass value is only 32.2 W/kg after Martian considerations, the extra mass might as well be spent on batteries or fuel cells without regeneration. The second alternative suggests implementing a land microwave or laser electromagnetic targeting source to beam energy towards the UAV while it is in Table 12: Memory allocation and data storage for distributed sensor package for the gyroscope and accelerometer data in order to resolve the analog output. Given the memory requirements listed in table 12 below, the Aeroflex UT28F256 LV PROM can store over twelve thousand data sets given a sample rate of once per second and 22 bytes per set. This allows for over two hours of storage time, which is adequate given the current mission specifications. The SEAKR FLASH chip is also able to store hundreds of images at any one time. In order to ensure a timely and accurate transfer of data to orbiting satellites, a first-infirst-out (FIFO) queue structure will be implemented in memory where old data will be transmitted first. 20 flight. The reason for the rejection of this alternative is simple – there is no guarantee that any form of a land station will be available for this purpose, while satellites are too far away to be able to transmit and pinpoint at these wavelengths accurately. The 0.3 cm-30 cm wavelengths for microwaves and hundreds of nanometer wavelengths for lasers implies that the energies cannot be precisely targeted from the orbit. We attempt using fuel cells in the third alternative. The fuel cell is an aspiring future technology that generates a very high energy per mass value. However, all existing fuel cell technologies require the use of hydrogen and oxygen for the generation of electric current. On the other hand, Mars is dominated by CO2, nitrogen, and argon. One solution to the paucity of hydrogen and oxygen is to bring along pressurized/refrigerated gas tanks; another solution is to produce oxygen via compressing CO2 in the manner of the Mars In-situ Propellant Production (MIP). As for now, the weight and size of a custom MIP device is not readily available, nor has it made for rapid collection of CO2 (only at night and low temperatures of around 200 K) although an 8.5 kg, 40cm x 24cm x 25cm device has already been demonstrated. A fuel cell system will require the following components to carry the necessary hydrogen and oxygen onboard the UAV: liquefaction. Moreover, pressurization requires an entire system of compressors with the addition of extra volume. The same volume concern holds if a refrigeration system is used. As the volume tradeoff shows, fuel cells without the best pressurization are not optimal for an aircraft of this size, even if the mass required for the energy supply is small. Until the Mars In-situ Propellant Production can provide immediate oxygen production, or until the volume can be drastically reduced by using absorption material or other Martian gases like nitrogen, fuel cells for the UAV will have to hold. Batteries: Batteries are by far the more convenient and readily available technology compared to fuel cells, and require no peripheral equipment at a much more compact volume. A variety of battery types are commercially available and under development, as listed in Table 13 below: Name Lead Acid NiMH Li+ NaS Li+ SPE kWh/kg 0.035 0.07 0.15 0.11 0.25 Table 13: Comparison of battery energy/mass ratios Tanks (hydrogen, oxygen) with insulation Fuel cell pressurization maintenance (or even refrigeration) system Fuel cell array and delivery Control system Mass and Containment Considerations: Lithium ion Solid Polymer Electrolyte (SPE) is the best contender among these existing technologies. It provides the highest energy density, around 1.5 to 2 times more than the currently existing Lithium-ion battery technology. The sample that we have chosen is under development by Ultralife since 2001, and is a feasible power source due to its nonatmospheric requirements. A simple calculation at an allowance of 4.2 kg gives 1050 Whr, while the volume of the battery is around 400 Wh/liter or 400 kWh/m3; 1050 Whr yields 0.0026 m3 of theoretical battery space, or around 14cm x 14cm x 14cm of volume. Compared to fuel cells, the given mass-volume tradeoff is extremely reasonable for our considerations. Sources show that 1g of hydrogen fills 11 liters at 0°C and 1 atm. In order to reduce the volume of the tank to reasonable levels, we can try to pressurize the volume to around 2.2 liters at 5 atm on a Martian surface pressure of 0.01 atm, which is a very generous pressurization value. If the temperature in the fuel tanks can be maintained at around -100°C, via simple insulation and without extra refrigeration cycles, the volume can fall to around 1.4 liters for 1 gram, or 0.7143 g/liter or 0.7143 kg/m3. Given that we have at most 0.008 m3, or a 20cm x 20cm x 20cm, volume available for a H2 tank, 5.7 mg of H2 is required. In this case, the tank will fill up 10% of the wingspan and nearly 82% of the maximum fuselage area. Now consider that 1 kg of H2 gives 86 MJ of energy, which means that multiplying 5.7 mg by 86 MJ/kg we will have 491 kJ, or around 137 Whr, which is still less than the best batteries (>200 Whr) at the cost of a much larger volume. This result only takes into consideration the hydrogen tank. The oxygen tank adds to another part of the fuselage, and is usually larger without Navigation: The absence of magnetic poles makes navigation on Mars particularly difficult without a GPSlike system. Since the flight of the UAV will be decided upon in advance and, in general, will be relatively direct across the surface of Mars, the UAV can take advantage of its on-board sensors such as accelerometers and gyroscopes to detect any deviations 21 from its path. Such sensors should be sufficient to ensure that the UAV stays on course once it has begun its flight path. However, for the UAV to begin on the correct flight path our design requires additional sensors. The photovoltaic compass (PV or sun compass), based on InGaP/GaAs/Ge technologies, can be used in the initial state of the plane’s launch to detect the proper orientation of the plane relative to the sun. Precise knowledge of the launch area on Mars will make it possible to know what the proper angle to the sun should be and this can be checked by the PV compass throughout the flight. The design of the PV compass is based on 26 small rectangular TripleJunction solar cells that are arranged in an octagonalcylinder fashion where the inward facing cells form the walls and bottom of the cylinder. The top surface is covered by anti-reflective fused Silica industrial-grade glass that has a low refraction coefficient relative to the Martian atmosphere, and low coefficient of thermal expansion. The glass part of the PV compass will be exposed to the Martian atmosphere at the top of the UAV’s fuselage without altering the laminar flow over the UAV. Incident sunlight will enter the PV compass and strike certain solar cells. Based on the current produced by each cell we can determine the angle of the incident sunlight and thus the angle to the sun. when the sun is directly overhead. This problem can be avoided ahead of time by taking this problem into consideration when planning the flight location and time. Another complication is that our measurements are accurate only when the UAV is flying parallel to the Martian surface. This means that measurements made during ascent/descent or turbulent flight should be discarded. Fortunately we will ensure that the PV compass is used only when the plane is in the proper orientation by checking the accelerometers and gyroscopes. Communication: In order to transmit the information gathered by the array of sensors aboard the UAV, we have included three UHF patch transceiver antennas in our design. Only two of these antennas will be operating at any one time, with the third antenna serving as a backup. We estimate the power requirement of two antennas to be 15 W with a total mass of 0.75 kg for the three antennas. One antenna will be one on each wing with the third on top of the fuselage. The patch antennas have a flat profile which can further be reduced by placing them into indentations in the surface structure. These particular antennas are compatible with existing satellite/rover communications equipment. Rather than attempt to transmit directly to Earth, we assumed that the UAV would be operating in an environment where there would be multiple opportunities to transmit data to Mars-orbit satellites. By avoiding direct transmission to Earth, we save power and thus weight aboard the UAV. Also, we reduce the probability of corrupting the transmitted signal. If the UAV could continuously transmit to a satellite, our communications uplink capability would be 8 MB/min per antenna. Implementing this design with only the current satellites orbiting Mars, the Odyssey and the Global Surveyor, would permit one 8 minute window during the flight to transmit all of our data, which would have to be limited to 64 MB. A future satellite communications infrastructure around Mars would increase the value of the UAV’s mission by allowing for the transmission of higher resolution images. Furthermore, this network of communications satellites could serve as the backbone of a navigational system for this and other missions to Mars. 90 mm 215mm 37 mm 37 mm 37 mm Software: Software will be written in a combination of C and Assembly based on industry-standard MCS-96 RTR instruction set architecture. External events will be monitored using multi-level state diagrams clocked at 20 MHz and eighteen separate interrupt service routines (ISRs) such as the timer COMPARE interrupt and UART RXC interrupt. A real-time microcontroller Figure 15: Photovoltaic Compass With this design, we can detect 26° of incidence across each solar cell and can determine the angle to the sun with high precision based off of partially lit cells. One shortcoming of the PV compass, however, is the ambiguity of direction that remains 22 operating system can be used to provide nearinstantaneous responses to external events. All software packages will conform to current IEEE Code of Ethics and ACM standards. except the ACT5108, which is instead located on a separate board near the DC engine. The sensor package is distributed throughout the UAV’s fuselage to provide maximum coverage. The batteries are regulated with a separate controller chip and a series of DC-DC converters to maintain the voltage and provide adequate Vcc for the various components. Block Diagram: Figure 16 shows the high-level block diagram of the electrical systems. The central microcontroller unit contains all the components listed in Table 11 LIDAR Software Antenna Txmitter DAC Rxver ADC Mechanical Transducers Hi-Res Camera ADC DSP Video MEM Card MCU DAC DSP DAC Left Wing DAC Txmitter ADC Rxver Mechanical Transducers Right Wing ADC Sensor Package Antenna Solar Compass ADC DAC Tail DC Motor Controller Controller Batteries DC Electric Engine Mechanical Transducers Propeller Copyright 2004, CU Odysseus Team Figure 16: Functional Block Diagram of Electrical System 23 that interprets data from accelerometers, gyroscopes, and a photovoltaic compass. Microcontrollers and the associated software will process sensor data to check for deviations from flight path and will implement the necessary corrective flight adjustments. For communication between the UAV and Mars-orbit satellites we have, for redundantly, chosen three UHF patch transceiver antennas that can continuously transmit data produced by the UAV instrumentation. Finally, a power supply consisting of multiple lithium-ion SPE batteries will be responsible for powering the instrumentation and propeller motor. Lightweight batteries are chosen upon detailed consideration of watt-hours per kg of all potential power sources including solar cells. Due to the constraints of time and expertise, our UAV design has excluded discussion of the induced roll due to a propeller. Furthermore, we have not specified an exact design for the contours of propeller itself. These elements of aircraft design are extremely complex and exceed the expertise of the team. Additionally, power supply imposes limitations on flight duration. In the future it is possible that potential power supplies, which we have also explored, will improve to a point where they are feasible for use on an unmanned Martian aircraft, extending its flight duration. Despite our omissions, we our confident that the team’s UAV design is an optimal solution with existing technologies and will provide a reliable platform for Martian exploration. Conclusion Throughout the design process, the Odysseus Team has striven to maximize the versatility of its UAV proposal. The team optimizes factors such as mass, power, speed, and flight duration to devise an aircraft that will meet the demands of a Martian scouting mission. The product of such systematic and collaborative design process is a 10.8 kg aircraft that can fly unaided by a human controller for 2.3 hours at 130 m/s and a cruising altitude of 500 meters with a maximum flight range exceeding 1000 km. Equipped with topographical and imaging instrumentation, such a UAV will be able to produce the detailed information necessary for future manned missions to Mars. The UAV blueprint outlined in this paper includes a 2.1-meter long fuselage, a 0.25-meter maximum diameter for the fuselage, and a 2.078meter wingspan. The design of such components takes into consideration low atmospheric density which causes reduced lift and drag. The dimension of wingspan is the result of detailed lift and airfoil analysis. To propel the UAV in an atmosphere that lacks the necessary amount of oxygen for combustion we have chosen a single 2.27-meter diameter propeller mounted aft of an inverted V-tail. The decision to use an inverted Vtail is the result of a compromise between the higher control stability but higher drag of a conventional tail and the lower drag but lower control stability of a V-tail. To control the UAV flight path we have designed a feedback system 24 References Aeroflex UTMC. (2002, December). UT28F256LV Radiation-Hardened 32Kx8 PROM. Aeroflex UTMC. (2003, May). UT80CRH196KDS Microcontroller Datasheet. Aeroflex UTMC. (2003, May). ACT5108 & ACT5109 3-Phase Brushless DC Motor Drivers. Aircraft Design & Aeroflightdyanmics Group (ADAG). Aircraft Designed at DPA. Retrieved February 20, 2004 from http://www.dpa.unina.it/adag/eng/aircraft-design.htm Anderson, J.D. (2000). Introduction to Flight (4th ed.). Boston, MA: McGraw-Hill Companies, Inc. Archer, R.D. (1996). Introdution to Aerospace Propulsion. Upper Saddle River, NJ: Prentice-Hall, Inc. Artemis Society International. (2001). Artemis Project: Hydrogen-Oxygen Fuel Cells for Lunar Habitat Energy Storage. Retrieved from http://www.asi.org/adb/04/03/03/h2power.html Ashby, M. & Cebon, D. (1996). Case Studies in Materials Selection. Retrieved October 2003 from http://www.analog.com. Beer F. & Johnston, R. (2001). Mechanics of Materials (3rd ed). New York, NY: McGraw-Hill. Borst, H.V. (1973). Summary of Propeller Design Procedures and Data (Vol. 1). Springfield, VA: National Technical Information Service. Cesarone, R., Hastrup, R., Bell, D., Lyons, D., & Nelson, K.. Architectural Design for a Mars Communication & Navigation Orbital Infrastructure, AAS 99-300. Cognex. (2001). MVS-8100 Datasheet. Cognex. (2002). CDC Series Digital Machine Vision Cameras.. http://www.cognex.com. Darooka, D.K. & Jensen D.W. Advanced Space Structure Concepts and Their Development, AIAA-2001-1257 http://www.ilcdover.com/WebDocs/AdvStct.pdf DSP Architectures. (2004, January). RHDSP24 Radiation Hardened Scalable DSP Chip. EADS Space Transportation. (2003). EADS Space Propulsion. Retrieved October 27, 2003 from http://cs.space.eads.net/sp/ Edie, D. & Buckley, J. (1993). Carbon-Carbon Materials and Composites. USA: William Andrew Publishing. EG&G Services Parsons, Inc. (2000, October). Fuel Cell Handbook (5th ed). Retrieved from http://www.fuelcells.org/fchandbook.pdf. Endevco, Inc.. (2003). Silicon MEMS Pressure Sensor, Model 32394. http://www.endevco.com/ Ezell, E.C., & Ezell, L.N. (1984). On Mars: Exploration of the Red Planet, 1958-1978. Washington, DC: NASA Hall, D., Tsai, K.C., Galbraith, D., & Parks, R.W. (1997). Conceptual Design of the Full-Scale Vehicle Propulsion System Concept. Airplane for Mars Exploration. NASA/Ames Research Center, CA: David Hall Consulting. Hepperle, M. (2004). Aerodynamics for Model Aircraft. Retrieved February 18, 2004 from http://www.mhaerotools.de/airfoils/index.htm 1 Hexweb. (2004). Hexweb: Honeycomb Attributes and Properties. Retrieved February 20, 2004 from http://www.hexcelcomposites.com/Markets/Products/Honeycomb/Hexweb_attrib/hw_p11.htm James, D. (2003). NASA Quest: Atmospheric Flight. Retrieved November 2003 from http://quest.arc.nasa.gov/aero/planetary/atmospheric/structure.html Kaplan, D.I. et al. (2000). In-Situ Propellant Production on Mars: The First Flight Demonstration. Retrieved from http://powerweb.grc.nasa.gov/pvsee/publications/mars/MIP_LPSC.html Kroo, I. & Shevell, R.S. (2003). Aircraft Design: Synthesis and Analysis. Stanford, CA: Desktop Aeronautics, Inc. Retrieved February 21, 2004 from http://adg.stanford.edu/aa241/AircraftDesign.html Lathrop, J., Snider, R., Becraft, W., Doll, D., & Cianciotto, F. UAV and LIDAR Mine Detection in the Surf Zone: Operational Considerations. General Atomics Aeronautical Systems, Inc. Mason, L. (2000). Prototype Solar Panel Development and Testing for a Mercury Orbiter Spacecraft. AIAA-2000-2881. Cleveland, OH: NASA Glenn Research Center. Morton, O. (1999, December). MarsAir: How to build the first extraterrestrial airplane. Air & Space Magazine, Vol. 14. Retrieved October 28, 2003 from http://quest.nasa.gov/aero/planetary/MarsAir.html Motorola, Inc.. (2004). http://www.motorola.com. Mueller, T.J. (1985). Low Reynolds Number Vehicles. AGARDograph, No. 288. Murray, J.E. & Tartabini, P.V. (2001). Development of a Mars Airplane Entry, Descent, and Flight Trajectory. Retrieved October 2003 from AIAA. Nice, K. (2003). How Fuel Cells Work. How Stuff Works. Retrieved October 20, 2003 from http://science.howstuffworks.com/fuel-cell.htm Nicolai, L.M. (1975). Fundamentals of Aircraft Design. Fairborn, OH: E.P. Domicone Printing Services. Ralph, E.L. & Linder, E.B. (1996, May 13). Advanced Solar Panel Designs. IEEE Xplore Journal. 25th PVSC. Washington D.C. Raymer, D.P. (1999). Aircraft Design: A Conceptual Approach (3rd ed.). Reston, VA: AIAA Education Series. SEAKR Engineering, Inc. (2002, October). NV-CPCI Non-Volatile Compact PCI Memory Card. Shevell, R.S. (1989). Fundamentals of Flight (2nd ed.). Upper Saddle River, NJ: Prentice-Hall, Inc. Stackhouse, D. (1999, January). Conventional vs V-Tails. Charles River Radio Controllers. Retrieved March 5, 2004 from http://www.charlesriverrc.org/articles/design/donstackhouse_conventionalvsvtail.htm Stanford Audio Phonic Photographic Infra Red Experiment. (1995, February 22). Space Environmental Study. Retrieved from http://students.cec.wustl.edu/~sapphire/design/log/spacenv.html Stevens, B.L., & Lewis F.L. (2003). Aircraft Control and Simulation (2nd ed.). Hoboken, NJ: John Wiley & Sons, Inc. Teofilo, V. L., Brown, R., & Cuellar, E.A. (2001). Advanced Lithium Ion Solid Polymer Electrolyte Battery Development. United States Department of Energy. (2004, January 15). Energy Savers: Electric Vehicle Batteries. Retrieved from http://www.eere.energy.gov/consumerinfo/factsheets/fa1.html. 2