Sei sulla pagina 1di 89

Comparative Analysis of Three Concepts for Aerostat Based Electrical Power Generation System

Gordon Civalier1[905288994], Christopher Fridley1[905107837], Jacques Li1[905269834], Justin Seabe1[905090891], Adriana-Xenia Stanciu1[905108747], Michael Willey1 [905091176]

and
Rajkumar Pant2

Virginia Polytechnic Institute and State University, Blacksburg, VA, 24061

19 April 2011

1 2

Senior Undergraduate Student, Aerospace & Ocean Engineering Department, Student Member AIAA Visiting Associate Professor, Aerospace & Ocean Engineering Department, Member AIAA

Table of Contents BACKGROUND AND INTRODUCTION.................................................................................................. 4 CURRENT AND PROPOSED WING ENERGY SOLUTIONS ................................................................. 4 Conventional Ground Turbines................................................................................................................. 4 Airborne Wind Energy Converters ........................................................................................................... 7 Vertical Axis Airborne Wind Turbines ................................................................................................ 7 Diffuser Augmented Wind Turbines (DAWTS) .................................................................................. 8 Savonius Type Wind Turbines ........................................................................................................... 10 CURRENT AND PROPOSED SOLAR ENERGY SOLUTIONS ............................................................ 10 Ground-based Solar Power Technology ................................................................................................. 10 Airborne Solar Energy Converters .......................................................................................................... 10 SCHEME FOR COMPARATIVE ANALYSIS ......................................................................................... 12 KEY DESIGN ISSUES AND CONSTRAINTS ........................................................................................ 12 DESCRIPTION OF THE THREE CONCEPTS UNDER INVESTIGATION .......................................... 13 AEROSTAT SYSTEM COMPONENTS ................................................................................................... 14 Envelope ................................................................................................................................................. 14 Material Properties .................................................................................................................................. 16 Fabrication of envelope sections ........................................................................................................ 18 Material Selection .............................................................................................................................. 18 Gas management ..................................................................................................................................... 19 Mooring Platform.................................................................................................................................... 20 Winching ............................................................................................................................................ 21 Tether ...................................................................................................................................................... 22 Tether Weight and Resistive Loss ...................................................................................................... 26 Effects of Altitude Variation .............................................................................................................. 27 Confluence Lines................................................................................................................................ 27 Drag due to Confluence Lines ............................................................................................................ 27 Number of Lines Required ................................................................................................................. 29 COST METHODOLOGY .......................................................................................................................... 30 General Manufacturing and Labor Assumptions .................................................................................... 31 General Winching Cost Assumptions ..................................................................................................... 31 Electricity Generation Sources for Comparison ..................................................................................... 32 SHROUDED TURBINE CONCEPT ANALYSIS ..................................................................................... 33 Shape Optimization................................................................................................................................. 33 Aerodyanamics of Annular Airfoils ........................................................................................................ 38

Shroud Geometric Parameters ................................................................................................................ 41 Aerodynamic Analysis ............................................................................................................................ 47 Stability Analysis of Annular Shroud ..................................................................................................... 50 Stability Analysis of Complete Aerostat ................................................................................................. 53 Material Stress Analysis ......................................................................................................................... 57 Diffuser Augmented Wind Turbine Cost ................................................................................................ 59 General Assumptions ......................................................................................................................... 59 Cost/Pricing Analysis ......................................................................................................................... 59 Cost of Electricity Generation ............................................................................................................ 60 Conclusions ........................................................................................................................................ 61 MAGENN MARS CONCEPT ANALYSIS ............................................................................................... 62 Scaling Methodology .............................................................................................................................. 62 Scaling Results ........................................................................................................................................ 63 Aerodynamic Analysis ............................................................................................................................ 64 Tether Profile and Mooring System ........................................................................................................ 67 Magenn MARS Concept Cost ................................................................................................................ 69 General Assumptions ......................................................................................................................... 69 Cost/Pricing Analysis ......................................................................................................................... 69 Cost of Electricity Generation ............................................................................................................ 70 Conclusions ........................................................................................................................................ 70 SOLAR AEROSTAT CONCEPT ANALYSIS .......................................................................................... 71 Preliminary Analysis ............................................................................................................................... 71 Methodology and Results for Typical Aerostat Shapes ..................................................................... 71 Revised Sizing ........................................................................................................................................ 73 Revised Envelope Sizing Methodology ............................................................................................. 73 Revised Envelope Sizing Results ....................................................................................................... 76 Mooring and Tether ................................................................................................................................ 79 Multi-tether analysis ........................................................................................................................... 79 Solar Aerostat Cost ................................................................................................................................. 80 General assumptions .......................................................................................................................... 80 Cost/Pricing Analysis ......................................................................................................................... 80 Cost of Electricity Generation ............................................................................................................ 82 Conclusions ........................................................................................................................................ 83 Solar Concept Conclusions ..................................................................................................................... 84 FINAL CONCLUSIONS ............................................................................................................................ 85
3

References .................................................................................................................................................. 86

BACKGROUND AND INTRODUCTION Environmental concerns over conventional power generation and rising energy costs have led to the need for alternative energy solutions. As the world enters a new century with an accelerating demand for energy, successful management of energy resources is vital to any countrys future. Current estimates predict world energy consumption will rise about 50% by 2035, the majority of increase occurring in developing countries (1). In the U.S., electrical energy production has relied heavily on coal for the past half century. In 2009, coal provided 48% of electrical power to the U.S (1). However, U.S. coal already reached peak production 8 years ago (2), and its consumption has declined in recent years as well. The U.S. must look to other sources of energy to fill the gap left by a falling supply of coal. The wind and solar power industries are expanding to fill that gap. From 2007 to 2009, wind energy generation more than doubled and photovoltaic solar generation increased 14-fold (1). The recent growth of these renewable sources has proven their viability despite having leveled costs much higher than coal. Their success can be partly attributed to a rise in public interest in renewable resources. Therefore, some margin of cost above coal, or current wind power, may be permissible for new power generation concepts as long as non-economic benefits are apparent. This study discusses the feasibility of electrical power generation by using a tethered aerostat and the marketability comparison among these concepts, and aims to carry out a techno-economic comparison of three such concepts that have been proposed. An aerostat based system is inherently capable of high wind energy extraction capability, and/or unhindered access to sunlight due to higher altitude of its deployment. However, the cost of producing electricity by such a system should be of the same order or better than that of conventional power generation systems attached to the national power grid, for it to offer a long-term viable alternative.

CURRENT AND PROPOSED WING ENERGY SOLUTIONS

Conventional Ground Turbines The current wind power industry is used as a benchmark for our design. The recent success of conventional turbines has proven their viability despite having a levelized cost 50% higher than coal. The success can be partly attributed to a rise in public interest in renewable resources. Therefore, some margin of cost above coal, or current wind power, may be permissible for an AWT concept as long as non-economic benefits are apparent. One of our team members recently attended a tour of Beech Ridge Wind Farm in Greenbrier County, West Virginia. Seeing the turbine farm first hand and consulting with the workers there provided us with a first-hand perspective of the industry. Considering the obstacles associated with constructing the wind turbine farm can be useful to consider in our design project. The Beech Ridge Wind Farm was first submitted for proposal in 2005, but it took four years until construction could begin due to opposition from local residents. Opponents voiced concerns including potential reductions in property values, adverse effects on tourism, bat fatalities, claimed inefficiency of wind energy, tax benefits for wind factories, environmental degradation, and stresses on the local
4

infrastructure. Beech Ridge has claimed that the development will create local jobs, add to the local tax base and satisfy the electricity demands of 50,000 households. After years of expensive litigation, construction was finally allowed on terms that turbines could not be operated under darkness due to the local population of an endangered species of bat. The power output of the farm is halved because of this. Generally, wind farm sites are chosen in remote locations to avoid the opposition seen at the Beech Ridge site. However, Beech Ridge is remote enough that it required the installation of a 14 mile long high-voltage transmission line and a network of mountain roads; and still a small group of local residents made a large impact on the operation. Local opposition is therefore a costly factor in wind turbine construction, whether it means delaying construction for litigation or installing expensive transmission lines and roads to reach a more remote site. Airborne wind turbines have the potential to assuage some of the concerns with conventional wind farms. The mooring platform for an AWT requires the ground space equivalent to a large truck, so land leasing may be cheap, easy and non-invasive. Because the wind turbine is lifted above the horizon line it will not be as visually invasive as ground turbines. The height of the turbine also reduces ground noise and shadowing, and avoids the flight zones of local wildlife. Workers at Beech Ridge cited cases where disgruntled locals actually shot firearms at wind turbine blades, causing thousands of dollars in damages. This would not be a problem at our design heights. AWTs have potential operational benefits over conventional turbines as well. An important design requirement for ground turbines is that the bearing system must withstand the cyclic loads, which result from a vertical velocity gradient along the length of the blades. Workers at Beech Ridge claimed that the wind speed at the top of the blade plane can be 15 mph larger than at the bottom. Our calculations show that a wind increase of 20% is expected from bottom to top of the blade plane for a GE 1.5 MW Series. (Note: the GE 1.5MW Series wind turbine will be used frequently for comparison because of its abundance in the industry). Comparatively, a speed increase of less than 1% is expected for the same turbine at 500-m height. Thus, the bearing systems on AWTs can be reduced in complexity, cost and weight. The established method of calculating wind speed in wind turbine engineering is to use an exponential variation in wind speed with height. According to this method, wind speed can be defined relative to wind measured at a reference 10-m height as: () ( ) ( ) where vw(h) refers to velocity of the wind at height, h [m/s], v10 [m/s] - velocity of the wind at height h10 = 10-m, and a is the Hellman exponent. This formula explores wind speeds from 80-m height to our design height of 500-m, assuming a Hellman exponent of 0.16 for flat land. For means of comparison it was considered that, according to the U.S. Department of Energy: areas with annual average wind speeds around 6.5 m/s and greater at 80-m height are generally considered to have suitable wind resource for wind development (3) If it is assumed that the same wind speed is desired for development of an AWT at 500-m height, the wind speed at 80-m height would only need to be about 4.8 m/s. This fact greatly increases the land area of potential sites for AWTs as compared to ground turbines. Figure 1 shows the average annual wind speeds in the United States at 80 meters altitude. The regions in orange define the lowest wind speed for ground turbine sites. According to our simple estimation, AWTs have the potential for use in regions colored yellow and light green. Thus, AWTs could expand the range of wind power in the U.S. to heavily populated regions on the west and east coasts, instead of being confined to remote mountainous regions and the Midwest. In Figure 2, a map of transmission lines in the U.S., demonstrates how this range expansion could simplify wind power transmission. The transmission line map demonstrates a significant issue for the wind power industry: regions of efficient power transmission do not coincide with regions of efficient wind power production. AWTs
5

could take advantage of efficient transmission lines without the need to construct expensive new lines, as was done at Beech Ridge. (4)

Figure 1: Annual average wind speed in the United States at 80 meters altitude (5)

Figure 2: Wind Resource and Transmission Lines (5)

Airborne Wind Energy Converters There are three principal types of airborne wind energy converters (AWECs), primarily divided by the lifting methods used. The first is a helicopter platform, which is a heavier than air application. Lift is generated using mechanically actuated rotor systems, as in a moving helicopter vehicle. The primary drawback to such a design is the energy required to run the engine and keep the platform aloft, and the associated mechanical complexity that can negatively affect the system reliability. The second is a kite system, which uses the power of the wind to keep it aloft, just as in the case of a recreational kite. However, such a system can only stay aloft when there is a sustained wind, as a kite is still heavier than air. The simplest, and by far the most widespread, platform is the lighter than air system, which uses the buoyant forces of lighter than air gases (helium, hydrogen, etc.) to generate lift that does not depend on wind conditions or engines. Such systems can stay aloft in both calm and heavy winds, and have very few moving parts, making for a very reliable system. The kite concept can also be combined with a lighter than air system to form a so-called helikite, as shown in Figure 3. This is a very commonly employed idea, since the aerodynamic lift from the kite allows the balloon to be sized smaller and more efficiently, and provides a method of stability augmentation. However, such devices have larger variance in stable altitude depending on wind due to the inclusion of the kite. The studies in this report consider lighter than air systems exclusively, due to the high reliability of buoyant lift. Also, both helicopters and kites tend to require complex and often expensive control systems, which can negatively affect the reliability of the system. Lighter than air systems tend to be passively stable, and thus do not often require complex controls. (6) (7)

Figure 3: Photograph of aerostat with helikite configuration (7)

Vertical Axis Airborne Wind Turbines A vertical axis wind turbine (VAWT) suspended beneath an aerostatic balloon was considered. It was quickly determined, however, that VAWTs are not ideal for our applications because of low efficiency and high weight. According to a leading wind turbine expert, Mick Sagrillo, compared to horizontal axis turbines, vertical axis turbines are less efficient, and it takes more materials and labor to make the things. It is for this reason, he says, that they are virtually non-existent in market applications. The main benefits of vertical turbines are the ability to mount the generator on the ground and the possibility of operation at slow, varying winds. Neither of these benefits relates to a high-altitude turbine. A big problem is that VAWTs experience cyclical loading on the blade spars as the blades encounter different flow conditions on either side of the turbine. This leads to fatigue failure unless there is added
7

structure in the spars and bearings. This explains why VAWTs tend to be heavy. These cyclical loads will most likely create instability if the VAWT is suspended beneath a balloon, and may cause the VAWT to swing back and forth. It was decided that this concept was not workable, and the design was abandoned. Diffuser Augmented Wind Turbines (DAWTS) Diffuser wind augmented wind turbines are of economic interest because of a potential for greater efficiency compared to traditional wind turbines. A diffuser augmented wind turbine (DAWT) features a duct or shroud which encircles a horizontal-axis turbine and flares outwards downwind of the blade plane. Altaeros concept is essentially a DAWT incorporated into an aerostat system by use of an inflatable shroud. According to Altaeros, the diffuser shape adds passive aerodynamic stability to the aerostat. This is the result of the flared geometry which makes the shroud similar to a shuttlecock in its ability to align with the wind. The Grumman Corporation pioneered the study of DAWTs in a report published in 1981. The report includes estimates on performance and cost of various DAWT designs. According to their estimates: The demonstrated level of power augmentation with relatively unsophisticated model DAWT configurations (diffuser exit to entrance area ratio of 2.75) exceeded by a factor of four the output of a bare turbine of equal diameter at the same wind speed. Forecasts of possible DAWT performance for more improved versions of the diffuser and wellmatched turbine/diffuser combinations indicated power augmentation factors of between 5 and 8. (8) Because power is proportional to the area of the blade plane, it can be inferred that a DAWT blade diameter is at least half that of a bare turbine with equal power (for a power augmentation factor of four). Thus, a significant size and weight reduction is possible when a shroud is added. Their results also claim that a power output of 150 kW is obtainable for a blade diameter of 11 meters at a wind speed of 24 mph (11 m/s). This power output is on the same order as our required power, so this estimate should be kept in mind when calculating the rated power of our design. With regards to cost, the report concludes that small (5-200 kW) DAWTs constructed from fiberglass composites would be most likely to compete with conventional turbines. It should be noted that cost of composites has dropped significantly in the thirty years since this research was performed. Therefore, it is likely that this statement about composites is even more valid at present time. Grummans findings suggest that a DAWT can have characteristics which are ideal for an AWT: size reduction and power augmentation with respect to bare turbines, and usage of lightweight composite materials. In the case of Altaeros design concept, the addition of the shroud serves a dual purpose: augmenting the flow, and lifting the turbine. Therefore, incorporating an inflatable shroud may be a free augmentation of power, as long as the implications of producing such a novel aerostat shape are not detrimental to the design. Because of the inherent constraints of an inflatable shroud on the shroud geometry, it should be considered a relatively unsophisticated DAWT. Furthermore, it should be a goal to match a diffuser exit to entrance area ratio of 2.75, as stated. (8) Vortec Energy Limited and The University of Auckland have been doing research in DAWTs for about a decade. Over the course of the research, a series of modifications were introduced to their design and each failed in meeting expectations. The research brings to light common mistakes with DAWT design. Their main difficulty was in maintaining attached flow through the diffuser, which was solved by introducing inlet slots to energize the boundary layer. It was found that the inlet slot required an elliptical nosecone to prevent flow reversal, which occurred with sharp-edged inlet slots. In the current design, a converging portion of the duct is added, which they believe increases flow augmentation. Their data on
8

power output is invalid because of blockage effects of the wind tunnel the design was tested in; so, unfortunately, it cannot be used for comparison. This study suggests that flow separation through the duct is an important consideration in design. The methods described in this paper should be employed to prevent separation when designing a DAWT. Methods like directing free stream flow through slots downstream of the turbine would be complicated by an inflatable shroud. (9) Sheila Widnall, a renowned Aerospace Researcher of the Massachusetts Institute of Technology (MIT), recently published a two-dimensional flow analysis through a ducted turbine. By applying an incompressible potential-flow vortex method, she calculated the flow characteristics through the duct. The results suggest that the power output and the drag for a DAWT can be varied by altering the magnitude of velocity decrease at the turbine. Widnall plotted power and drag as a function of a nondimensionless velocity change, , in Figure 4.

Figure 4: Power and drag as a function of dimensionless velocity change

The drag plot incorporates the concept of leading edge suction, a negative drag component on the shroud due to a velocity increase upstream of the turbine blades. At large s, the shroud drag is actually reduced to zero. Note that this does not mean there is no longitudinal force on the DAWT; but that all of the force is on the turbine blades, not the shroud. Also note that this analysis does not consider viscous drag force. The power plot compares different exit-to-entrance area ratios, AR, to a bare turbine, labeled Betz. Notable is the fact that maximum power occurs at a lower when a shroud is introduced. The magnitudes of drag and power amplification factors seen in these results are likely inaccurate because 3-D and viscous effects are not considered. Empirical data from studies like Grummans should be used for estimations. However, general conclusions from Widnalls results, such as leading edge suction effects, should be considered. Widnalls report implies that DAWTs have drag characteristics which may be beneficial to AWT design. The inflatable shroud of Altaeros design could have low drag as compared to an equivalent aerostat without the leading edge suction effects of the turbine. Drag on an aerostat increases blow-by, the displacement in the direction of the flow, which necessitates additional tether length and weight. Thus, low drag is an important design goal to minimize the total weight of the aerostat system. (10)

Savonius Type Wind Turbines Currently, designs of an airborne wind turbine for power generation do exist. One such company investigating such technologies is Magenn; they are currently in pre-production phase of their airborne rotor system MARS. The design features a savonius type rotor wheel inflated with a lighter than air gas that is deployed at about 900ft in altitude. The proposed design power production rating would be 50kW, measuring 100ft in diameter, and weighing about 26,000 lbs. Magenn has come a long way since their first prototype MARS Alpha flew in April 2008. At that time, the system produced a maximum of less than 10kW of power and had a power efficiency of less than 2%. The prototype was mainly a technology demonstrator that proved that aerostat for power generation was indeed possible. (11) Through extensive wind tunnel tests, the Alpha prototype was greatly improved in design to allow a power efficiency of up to 33% and produce up to 55kW of electricity. The current model, as shown in Figure 4, features an efficient three-blade savonius type rotor coupled to generators on each side of the rotating axle. To reduce cost, the generators used will be acquired from the hybrid automotive industry. The system uses helium as the lifting gas, but a step over to hydrogen may take place in the near future as the helium supply industry is very volatile. Hydrogen is very attractive because it is readily available and relatively cheap. Concerns arise over hydrogens high flammability, but this may be mitigated in application such as off-shore deployment. Dihedral stabilizers situated on the sides of the rotor will ensure the inherent stability of the system. For the highest power output, it is necessary for the rotor to be facing into the wind at all times. With the stabilizers, this can be easily achieved without the use of complex active control surfaces. The envelop material is a readily available nylon base with polyurethane and includes a gas barrier to lower the gas leak rates. In addition, a UV protection surface will cover the external surface to increase the longevity of the fabric. The life expectancy of the fabric is about 7 years, while the structural life expectancy is about twice that. The system should be able to be self-sustainable for at least six month periods, with overhauls being done when the envelope reaches the end of its life. (12)

CURRENT AND PROPOSED SOLAR ENERGY SOLUTIONS Ground-based Solar Power Technology The current solar energy can be harvested for electricity either directly using photovoltaic (PV) cells or through concentrated solar power (CSP). CSP utilizes a system of mirrors to focus the suns thermal energy which is then converted into heat which drives a steam turbine connected to an electrical generator. Solar power systems range in size from personal household use to utility scale power plants. Although solar cell systems offer the benefit of low operational costs, like turbine systems, a solar cell system is very expensive due to the upfront cost of the silicon cells. Also, geographical location plays a big role in the cost effectiveness of implementing solar cells since it determines the amount of solar insolation that is available to produce electricity. Other negative impacts include; installation costs, proper direction, shadowing affects and periods of sunlight loss. Airborne solar power allows unobstructed view of the sun as well as higher solar insolation, which increases the amount of power production potential (13) Airborne Solar Energy Converters There are several designs currently in development utilizing lighter than air platforms to generate power from solar radiation. Most practical methods operate at low altitudes of less than 100 meters, and tend to employ oblong shaped balloons that maximize the surface area exposed to solar radiation. One
10

such method is the SunHope solar balloon, developed by architect Joseph Cory of Geotectura and aerospace engineer Dr. Pini Gurfil of Technion. The concept is very simple, involving only helium balloons constructed from photovoltaic cell coated fabrics at low altitudes, with the primary purpose of elevating balloons being to clear obstacles on the ground. According to Dr. Gurfil, a balloon with a 10 foot diameter with about 25 m2 of solar paneling can produce approximately 1kW of electricity. His analysis indicates an initial installation cost of about $4000 per 1kW balloon; in comparison, a ground solar field with the same power output has an initial cost of around $10000. These balloons would be ideal for off the grid applications, such as individual homes and third world countries, due to their simple design and tether system. A conceptual image of the design is shown in Figure 5. (14)

Figure 5: SunHope solar balloon concept (14) Dr. Eric Cummings Cool Earth solar balloon array method is a more sophisticated approach to the problem. Each array consists of several 2 meter diameter, oblate spheroid shaped balloons that are referred to as solar concentrators, an example of which is shown in Figure 6. Solar radiation enters the balloon through the clear top film and strikes the reflective bottom film. The bottom film is essentially a concave mirror, and thus has a defined focal point; the photovoltaic cell is placed at this focal point. The solar radiation is thus focused onto the photovoltaic cell, significantly reducing the area of photovoltaic cells required in the system and increasing the efficiency. The chief benefits of such a system are its high efficiency, and its ability to protect the mirror surface from weather and the environment. Cool Earth claims to be able to produce about 500W of electricity with each balloon, and they plan to reduce the production cost of each balloon to about $2. If this goal can be achieved, a 1MW installation of these balloons would have a per watt cost of about $0.29, compared to the nearly $4 per watt cost of similar installations using current technology. (15)

11

Figure 6: Cool Earth solar balloon concept (15)

SCHEME FOR COMPARATIVE ANALYSIS Each of the three concepts will be investigated individually to determine the marketable power output the system should be designed to produce. The optimal power output of the system is determined by the lowest lifecycle cost per kilowatt-hour. It is also important to note that the output rating should also be economically competitive with existing energy solutions. In achieving the goal of lowest lifecycle cost, other factors such as reliability, long lifespan, and lower initial cost are implied. Since the constraints and conditions vary between the three concepts, the optimal power output can differ for each of the three concepts. Therefore, it is difficult to justly compare all three designs against each other. Different applications will determine the best concept to be deployed in that environment and the amount of power required from such system.

KEY DESIGN ISSUES AND CONSTRAINTS An aerostat for power generation must be technologically innovative to be economically feasible in competition with conventional power production technologies. There are a few key issues and constraints that must be addressed and investigated to ensure the aerostats viability. A power generating aerostat should maintain a period of autonomy where no human interaction is required. This duration can be varied based on its inherent stability, lifting gas type, and permeability of the shroud material. The time aloft determines the maximum time allowed without the need for a maintenance crew to replenish the lifting gas. This duration can be further extended if an automatic gas replenishing system is installed. The longevity of such system should compare favorably with existing alternative energy systems. This can be achieved by the use of durable materials, and ensuring proper maintenance procedures. Increasing the lifespan of the system will in turn lower the cost of electricity produced, which will allow the aerostat to be economically competitive with established energy production technologies. Each component on the shroud has a different lifespan associated, so the system should be able to be overhauled in independent sections. The shroud envelope should be able to be replaced every few years while tether and generator can be serviced at different intervals. The aerostat is inevitably subjected to natural forces such as storms and high winds. The structure must be able to endure the elements by design of a sturdy shroud and a thick and strong tether. This
12

includes the appropriate envelope material and internal ribs. Aerodynamic stability, either inherent or added, is crucial for an aerostat in order to withstand inclement weather. The aerostat system should also allow for a certain level of mobility for applications such as military use, disaster relief, or remote location use. This can be achieved by tethering the system to a movable mooring system, such as a truck. The mooring and winching system also acts as an altitude control system that can adjust the tether based on percentage of excess lift, weather conditions, and conditions such as temperature and elevation. The constraint faced by any aerostat concept is the operation altitude. For a wind-based aerostat, the certain advantage of higher altitude is higher sustained wind speeds. However, air traffic control organizations such as the Federal Aviation Administration (FAA) dictate the upper ceiling of which a tethered aerostat can normally operate. This regulation is for the concern of air traffic safety, which must be carefully followed. A special permission would likely be required for aerostat operations above 500ft in altitude.

DESCRIPTION OF THE THREE CONCEPTS UNDER INVESTIGATION This report focuses on three different power generation schemes based on an aerostat platform. The first is an airborne wind turbine (AWT), held aloft by a lighter-than-air shroud, with a geometry similar to that shown in Figure 7. The concept is currently endorsed by Altaeros Energies, who seek to develop a commercially viable model based on the concept. The shroud is shaped such that the region of the flow behind the turbine blades serves as a diffuser; such shaping enhances the power of the turbine by significant margins. Such a system is generally termed a diffuser augmented wind turbine (DAWT). Studies on a 100 kW scale concept suggest that a per kWh energy cost of around $0.06 could be achieved, a 25% savings on average ground based wind power costs (8) (10). However, the necessarily larger geometry of such a system may prove prohibitive when scaled up to a utility scale system. The second concept is also a wind turbine based system, but the orientation is different. The configuration is termed a Savonius type wind turbine, and relies on the drag on the surface of the aerostat to rotate the assembly, as in Figure 7. Magenn Power, Incorporated has developed a working prototype based on this concept with a design power output of 30 kW, with the prototype generating 9.8 kW of power. The purpose of this study is to investigate the effects of scaling up the system will have on the technological and economic feasibility of such system. (11) The final concept utilizes photovoltaic cells on a lighter than air platform, with the goal of elevating the cells above ground level obstructions such as buildings and trees. Balloons with an oblate spheroid shape, like those shown in Figure 7, are ideal for this application, as their flattened shape allows the cells on the surface to capture more sunlight during peak sunlight hours and yet is still a fairly efficient shape for generating buoyant lift. A chief merit to such a design is its sheer simplicity, as unlike a wind turbine system, solar power generation requires very few (if any) moving parts. However, the spheroid shape has no passive stability in winds, as wind turbines do, and thus keeping the system aerodynamically under control may involve innovative tether concepts since compromising some of the photovoltaic surface in order to add stability surfaces is not an option (14).

13

DAWT LTA aerostat platform concept (16)

Magenn MARS Savonius type LTA airborne wind turbine (11)

Solar cell clad LTA aerostat system (17)

Figure 7: Summary of concepts for aerostat based power generation systems A feasibility analysis is carried out for two theoretical concepts (solar power and DAWT power), which are then compared to the existing prototype developed by Magenn, based on marketability and lifecycle cost.

AEROSTAT SYSTEM COMPONENTS Envelope A major component of aerostat designs center on the effectiveness of its envelope. Accordingly, the design of its material is instrumental to provide necessary durability, adequate strength, and an inherent low structural weight. Although the application of the DAWT for aerostat technology alters the envelopes shape, basic envelope construction is constant. Therefore, past studies (18) into the development of envelope technology shows that envelope material selection contributes approximately 38% to the empty weight of an airship. Subsequently, the expected empty weight percentage is expected to increase for aerostats due largely to reduced internal structural weight. Moreover, an envelope in aerostat design is part of the system that provides the primary lifting forces. Coupled with large structural weight concentrations, material selection can thus improve or reduce payload capabilities. Hence, envelope requirements are provided formally here: 1. Low gas permeability 2. Low structural weight 3. High load bearing capacity 4. Weather durability 5. High operation range 6. Simplistic maintainability With this is mind, such a diverse range of requirements makes envelope design a balance of multiple materials and layers. For this reason, an explorative comparison of previous envelope

14

construction for typical aerostat designs was performed. A baseline design (19) for envelope construction makes use of two layers joined by an adhesive shown in Figure 8.

Figure 8: Baseline envelope construction (19) Note each layer provides a specific function. Therefore, it is explained that the outer polyester film provides an environmental layer whereas the inner polyester fabric provides for both permeability and structural requirements. Unfortunately, designs of this nature were meant for short term, low impact missions. For that reason, research was further invested into an envelope design that better satisfies the DAWT aerostat requirements. It was found that typical envelope designs that met the various requirements consisted of an additional layer dedicated to gas retention. Hence, the ordering and purposes of each layer is provided by Figure 9 (20). With the addition of a separate layer, each material is thus allowed to cater to specific requirement. In return, overall performance is improved and optimization of a desired requirement is made possible.

Figure 9: Typical envelope construction for endurance (21) The study also delved into the weight breakdown of the various layers for this particular envelope design. Typical weight breakup of envelope designs is therefore shown in Table 1. It is shown that though the load bearing material provides a major component of the weight with 43%, the environmental coating is a close second with 39%. Thereupon, the selection of these two materials will dramatically affect the payload capabilities.

15

Table 1: Weight breakdown of various layers for design of envelope fabrics


Layer Environmental Coating Load Bearing Gas Retention Total Variation Specific Weight (g/m^2) 92 100 42 234 20 Specific Weight (oz/yd^2) 2.71 2.95 1.24 6.9 0.59 % of Total Weight 39 43 18 100 8

If the aerostat design was instead changed to a solar type array, additional layers become necessary. For example, a schematic of the envelope selection for the High Altitude Long Endurance (HALE) Airship (21) is provided in Figure 10.

Figure 10: HALE Airship envelope design (21)

Material Properties Attention is then directed to the materials available that could possibly meet the purposes of individual layers. In the field of load bearing materials, emphasis is placed on low weight while maximizing tensile strength. Gas retention is optimized with materials of low permeability rates. Lastly, environmental fabrics have particular resistance to moisture and UV light. Overall, two types of materials were then compared in order to find the benefits of each, woven or non-woven fabrics. A comparison (22) of the two fabrics show, non-woven fabrics have numerous advantages over woven fabrics. For example, woven materials suffer from manufacturability issues that do not allow the weight and thickness optimization of non-woven fabrics. Moreover, woven fabrics suffer from crimp illustrated in Figure 11. Crimp is characterized by tensile loading of woven fabric induced by transverse loads at fiber overlap sections. In consequence, fabric strength is reduced because of crimp which also leads to decreases in long-term fatigue and creep rupture performance.

16

Figure 11: Diagram illustrating crimp (22) Accordingly, Table 2 provides a comparison of a range of new materials developed by Cubic Tech Corporation (CT) to those woven materials that have been used in past envelope designs.

Table 2: Fabric materials for envelope design


Weight (oz/yd^2) *Tedlar PVF Films Silicone Coated Woven Nylon Mylar CT1.5HB UHMWPE Composite CT0.3HB UHMWPE Composite Medium Weight Flexible Composites PVC coated fabric CT35HB Vectran Composite CT35HB Aramid Composite Heavy-Weight Flexible Composites CT155HB UHMWPE Composite CT135HB PBO Composite 1.03 1.13 1.62 1.20 0.50 4.16 4.10 4.70 13.00 10.30 Weight (g/m^2) 34.84 38.31 55.00 40.69 16.95 141.10 139.01 159.36 440.77 349.23 Tensile Strength (lbf/in) 145.04 47.00 24.07 87.00 41.00 __ 523.00 587.00 2167.00 1934.00 Slit Tear Strength (lbf) 0.233 14.00 __ 108.00 38.00 __ 205.00 131.00 313.00 250.00 Permeability L/m^2/day 2.32 0.00 0.30 <0.2 <0.2 3.75 <0.2 <0.2 <0.2 <0.2

Material Light-Weight Flexible Composites

In addition, the slit tear strength is introduced as a comparative measure of a materials performance. The tear strength is the force per unit thickness required to tear a material. Test specimens are placed in a tensile test machine in which the specimen has a slit cut into it and the two ends are mounted in jaws that are moved apart at a fixed rate. The jaws are set to separate at a rate of 20 in/min (500 mm/min). For bonding of multiple layers to form a single envelope fabric, adhesives become a key component. Because of the high strength required, attention is turned to an epoxy for bonding purposes. An epoxy (22), used chiefly in strong adhesives, is a class of resins derived by polymerization from epoxides that cures when mixed with a catalyzing agent or 'hardener'. Epoxies are further divided into 2groups; 1-component and 2-component. A 1-component epoxy is aliquid isocyanate-terminated polyurethane whereas 2-component epoxies consists of the components: polyol and isocyanate. When the two components are mixed they form urethane groups. In comparison, a 1-component epoxy typically has a relatively higher molecular weight and reduced strength. Thus, to improve functionality a 2compontent epoxy is desired.
17

Fabrication of envelope sections The final design of the envelope shape will inherently be very large and its shape must be created by the joining of smaller sections. Thus, it became necessary to find appropriate methods of joining the envelope fabric without weakening its properties. Different processes along with brief descriptions of their methodology are accordingly listed (21): 1. Hot air welding: A continuous jet of hot air is directed in between the tape and the two pieces of fabric. The melted surfaces are then bonded under the pressure of two rubber rollers. This is the main process of seam welding. 2. Impulse welding: This process uses electrical resistance to heat up a plate. 3. Wedge welding: Similar process to hot air but uses a hot wedge as the heat source. 4. Radio Frequency (RF) sealing: Heat energy generated from high frequency waves is used to melt the inner layers of the fabric, which act as a resistance in the circuit. On melting, the fabric layers fuse completely and an air tight sealing occurs. For selection of a fabrication process, care must be taken to ensure ease of machine-ability. As difficulty of machine-ability increases, so do expenses. A balance of cost to quality must be found that will ensure an optimized design. Material Selection The envelope design selected for the Magenn and DAWT concepts involved three main layers previously shown in Figure 9. Additionally, selection of RF sealing allowed each design to balance the coast of machine-ability with an acceptable seam quality. Due to high force winds, the medium weight CT35HB Vectran composite was selected for structural layers. In comparison to the Aramid material, the Vectran composite was lighter, by approximately 0.0203 kg/m2 (0.6 oz/yd2). Note that this became necessary when the shear size of the aerostat was visualized. Thus, this small difference amounted to key weight savings later. Additionally, the Vectran composite provided nearly the same tensile strength at 91.6 kN/m (523.00 lbf/in) to the Aramid composite of 102.8 kN/m (587 lbf/in). Both of these composites provide excellent permeability rates, whereas the PVC coated fabric was completely ignored for its poor permeability. For gas retention, the key concept is excellent permeability rates while maintaining a low weight. Because the gas retention layer isnt specified for strength, the light weight CT0.3HB UHMWPE Composite was selected for its low permeability rate, and sizeable weight difference. At 0.017 kg/m2 (0.5 oz/yd2) it is the lightest material found that can deliver a permeability rate less than 0.2 L/m^2 over a 24 hour period. The outer environmental layer should be able to withstand wind and UV damage. Since weight was essential to every aerostat design, the light weight Tedlar PVF Film (23) was selected. It provided a light weight layer that withstood more damage than any other light weight material and was reasonably close to the performance capabilities of some heavier fabrics, such as the CT35HB Aramid Composite. A coat of urethane was also applied in order to improve slit-tear properties. Similar to the HALE airship, an envelope design for solar cell technology would involve one additional main layer. Following a philosophy of high efficiency, the solar cell aerostat would use the triple junction with monolithic diode, which have a mass per area of 0.85 kg/m2 and a thickness about 1mm. Since the main purpose of the solar cells is to absorb UV rays, a coat of urethane is not needed.

18

Lastly, the adhesive used for bonding was recommended by the technical report Analysis and Design of Robust Helium Aerostats (24) as Loctites Hysol E-20HP with a tensile strength 39.9 MPa and a shear strength 28.6 MPa when bonded to epoxy. Gas management An important consideration for any lighter than air vehicle, particularly for those with long flight times, is the control of vehicles buoyancy. Since the buoyancy of the vehicle is a direct result of the lifting gas used, this essentially reduces to a management of the lifting gas. The first and most basic consideration is the choice of lifting gas. The two most common lifting gases for lighter than air applications are helium and hydrogen, since they are the two lightest gases on the periodic table. Table 3 summarizes the important lifting characteristics of both gases.

Table 3: Important Lifting Characteristics of Hydrogen and Helium (6)


Gas Hydrogen Density at STP (kg/m3) 0.0899 Lift capacity at STP (kg/m3) 1.135 Considerations The lightest gas available, and thus has the highest capacity for lifting. Very flammable and is prone to leakage due to its very small particles. Abundant and cheap to produce. Inert gas with a lifting capacity only slightly less lifting capacity than hydrogen. Also prone to leakage, but not quite as much as hydrogen. Expensive to produce.

Helium

0.1664

1.058

For aerostats operating above altitudes of about 500 meters, a device called a ballonet is usually necessary. A ballonet is essentially an air bag integrated into the larger envelope that is inflated or deflated as the altitude changes. At ground level, the ballonet is inflated to a point where the aerostat has zero disposable lift. Pumps are then used to throw out air and allow the lifting gas to expand, increasing the buoyant force on the aerostat and thereby increasing its altitude, as shown in Figure 12. Thus, they are very useful in aerostats that must cover a large altitude range.

Figure 12: Typical ballonet system (25)

19

However, a ballonet system adds a significant amount of weight to the system, and thus is not practical when the vehicle operates at lower altitudes. For lower altitude aerostats, a passive system consisting of folded fabric restrained by elastic bands will typically suffice. This will allow the volume of lifting gas to expand and contract freely, rather than depending on pumps as in a ballonet (6). Elastic strips can also be used in place of ballonets. The main benefit is that it effectively reduces weight and manufacturing complications. Note that normally elastic strips are used for aerostat systems with low operational altitudes, less than 500 m. For greater altitudes than 500 m, the strips must be altered or use of ballonet is desired. All systems discussed in this study will operate below the 500m altitude, thus the elastic strips are chosen for each design. Elastic strips in principle allow the system to adjust altitude by controlling the maximum volume available. Each strip is integrated into the envelope seams. The area of the strips both in length and width is designed based on the amount of expansion that takes place from the ground level to the design altitude. Room for expansion is also provided for the diurnal temperature variation. Expressing the relationship between different states of gas for a given mass quantity is relatively simple using the ideal gas law. The ideal gas law is modified to: ( ) Knowing the volume differential, the required dimensions at maximum diameter of the envelope is calculated through the operational volume. Accordingly, even if a constant overpressure is desired such as the 15 %, than the volume at an operational altitude is: [ ] ( )

Although it is possible to individually determine strip dimensions, this study is concerned mainly with its cost addition. Therefore, in a previous study conducted by Indias Snow and Avalanche Study Establishment (SASE), it was found the percentage mass of elastic strips to the overall envelope mass is approximately 8% (26). Consequently, for purposes of cost estimation the envelope mass is adjusted appropriately.

Mooring Platform The mooring system is an important design driver since its primary function is to keep the aerostat stationary within a limited air space. In order to maximize the energy output, adjustments in location may be necessary at times depending on the source of energy (sun, wind). Thus, a mobile mooring station is most advantageous. The mooring system whether mobile of fixed -- needs to withstand any other forces generated by the environment like sudden gusts, floods, etc. The size and structure of the mooring station depends on the loading conditions implied by the aerostat. Several existing airship mooring masts can be considered and adapted for aerostats. A current mooring system used by Lockheed Martin is shown in Figure 13. This systems capabilities include: rotating turntable, a winch, support mast for transport (and maintenance), anchorage, mobility. The mooring mast includes only a support on a mobile vehicle for transport and maintenance and retractable stabilizers on the sides of the vehicle that can be deployed when stationary.

20

When designing the mooring system for this project it would be preferable to follow the capabilities of the design in Figure 13. In addition to the platform shown this system needs a winch that conducts power to a battery or transformer, as well as withstanding the high loads generated by the aerostat and environment. The mooring system can have a winching system to be able to retract the aerostat using the tether for position assistance and power conduction. Additional equipment for power management found on the station may be: charge controllers, storage batteries, inverters for output voltages etc. Also the mooring station may be able to service the aerostat for damages and gas-loss.

Figure 13: Lockheed Martin aerostat mooring masts (27) (28) When multiple tethers are needed to improve an aerostats stability, a swivel type mooring system like in Figure 14 can be implemented to provide anchorage to the ground.

Figure 14: Swivel Type Steel Mooring (29)

Winching The winching components goal is to provide enough tension and support while being easily retractable. It serves as a cable attachment point to the aerostat. An electric winch is a good candidate in this case. Its motor must be able to either maintain tension or to maintain a certain cable under non-loadbearing conditions. A standard winching system mounted on a rotatable platform that is capable of withstanding heavy tension and loading can meet the requirements for the wind based aerostats. The main tether needs to be attached to the winch shaft. A conductive metal inside the shaft of the winch can transfer the power

21

to a power-out cable towards a battery. The shaft however, must be properly shielded or composed of non-ferrous materials, to minimize the effects of the magnetic field that the coiled cable may generate. Tether One or more cables called the "tethers" anchors any aerostat and maintains the aerostat in its operation envelope. The tether for aerostat systems must withstand great tension, conduct electricity (grounding and power) while being lightweight. It also has to provide grounding to prevent electric arcs from forming on the ground generated as a result of the static electricity associated with high altitudes and thunderstorms. The tether may contain a light conduction shield to assist in the reduction of electrical interference from nearby sources (aircraft, power lines). Although, the main tether is usually a power conducting cable that needs to withstand the load applied, there have been studies investigating three and six-tethered concepts which aim to improve the stability of an aerostat (30) Research shows that high-molecular-weight polyethylene and carbon fiber meet the high strength-to-weight ratio criteria as materials for the tether component. Synthetic fibers like Dyneema and Spectra have yield strengths up to 2.4 GPa, at a density of 970 kg/m3 ; initial calculations were conducted for a 5.03E-5 m2 cross sectional area (31) Carbon fiber sleeves have a density of 1800 kg/m3 (32) per approximately 0.0033 m thickness (33). The components of a small diameter tether cable for maximum power transfer (minimum diameter) were found to be as shown in the tether components of Table 6 (34). The power conductive cables are aluminum 1100 alloy, due to its light weight and malleability. A standard weight value is shown first for such a tether, which can be compared to the ADRDE tether weight of 80 kg/km, used for aerostats (6). One can optimize the design by introducing Dyneema SK75/76, or Carbon -fiber as the strength member in the tether, however the ability of these to stretch, can be detrimental to the Al conductors. The results of an initial analysis of an optimum power transfer tether design (Table 4) is later used in the selection of the power conducting cable of each system.

Table 4: Weight of tether as a function of the type of reinforcing fiber % Tether Standard D. SK75 Carbon Fiber
Tether Components Axial Filler Rod 3 Power Conductors Core Void Filler Core Binding Tape Core Jacket Strength Member Weather Proof Jacket 3 Optical Fibers Final Weight (kg/km) % 0.3 32.4 10 0.8 4.5 22 22.6 7.4 0.3 33.1 10.2 0.8 4.6 22.5 23.4 7.4 102.3 0.3 33.1 10.2 0.8 4.6 109.71 23.4 7.4 189.51 (kg/km) 0.3 33.1 10.2 0.8 4.6 40.5 23.4 7.4 120.3

In Table 4, based on the maximum power transfer-minimum diameter concept, does not take grounding into consideration. It can be observed however, that using Dyneema SK75/76 yields an initial tether weight of 189.098 kg/km, while Carbon fiber is lighter but has lower yield strength according to Figure 15, which shows the free breaking length (km) at which a rope breaks under its own weight. It should also be noted that the fiber-optic cables are implemented as components of the tether since they be
22

used for communication purposes, and should be considered in the event that the aerostat becomes multipurpose. Permissible blow-by is accounted for in the fact that 30% extra tether cable will be included, while confluence lines are approximated to weigh 1% of the envelope weight.

Figure 15: Free breaking length for rope materials (values in km) (31)

Taking into account the need for grounding and the complexity of the tether, separate grounding cables were considered for the purpose of conducting static electricity. Even though the cable would not need to hold any load the risk that fractures could occur for Al 1100 due to blow-by wind was determined too high. Any attempt at incorporating grounding into the main tether seemed to result in either tether failure after a limited number or lightning strikes, or mooring station and equipment damage (35) (36). Additionally, the aluminum-steel grounding cable would need a light anticorrosive coating that will not add significant extra weight. As a result, it was decided to implement a tether that can be found on the market already, and has been used for other power generating aerostats. Additionally, the power conducting cable need not necessarily withstand loading if multiple tethers are needed for stability of an aerostat. Dynema SK tethers can keep the aerostat moored, due to its properties mentioned in the earlier analysis. Another winching system is needed for the non-load-bearing conducting cable to adjust the altitude of the cable to be greater than that of the main tethers as to sustain no loading. The conducting cable decided upon is the tether design already found on the marked that the Magenn prototype - which is a high torque and produces low voltages uses. For a given amount of power, a low voltage requires a higher current and vice versa. The following tether design is needed to handle appropriate power loads by all three power generating concepts. Magenn has provided a specifications for their 0.035 m diameter tether as presented in Table 5 (37) . The length of the tether varies for every concept depending on the operation and deployment of each.

23

Table 5: Magenn Tether Weight Contributions


Component Copper Conductor Insulation Jacket Grounding Shield Total: Quantity 0.51 0.06 0.36 0.07 1.00 Units kg/m kg/m kg/m kg/m

A method for estimating the tether profile was developed from a computer program by Wright (38). The profile determination starts at the top of the tether with a short straight segment. The angle and tension in this first segment is determined from the excess lift and drag of the aerostat. The angle and tension in the next lower segment are then calculated from the weight and drag on the above segment, using the geometry shown in Figure 16.

Figure 16: Tether profile geometry The drag on each segment can then be found by estimating the Reynolds number at the altitude of the segment, and then calculating the force based on drag coefficients of a cylinder. This sequence continues until the tether reaches the ground, the tension reaches zero, or the tether becomes horizontal. The latter two results mean that the aerostat fails at providing enough force to lift the tether. Examples of tether profiles for a successful tether profile and failing profile are shown to demonstrate the method:

24

Figure 17: Successful and unsuccessful tether profiles The tether profile in the sizing estimation can be plotted based on the sizing parameters. The Inputs for the tether profile method include: excess lift determined from sizing (approximately 370 kg-f), a velocity profile with the design velocity (8 m/s) at 500-m, and tether specifications taken for the Standard tether from Table 4: Weight of tether as a function of the type of reinforcing fiber. The drag force on the aerostat is the remaining input, which must be inferred based on the assumption of a length 30% greater than design height of 500-m, or 650-m. The drag force was determined to be about 277 kg-f to obtain a tether length of 650-m. It should be noted that in this conceptual analysis, the actual design drag on the aerostat is unknown. Therefore, the variation of tether length with the aerostat drag is useful to consider for future drag estimations. For the same parameters, the percent extra tether (with respect to tether length with zero drag) is plotted as a function of aerostat drag:
Extra Tether Required vs. Aerostat Drag 90 80 70

Percent Extra Tether, %

60 50 40 30 20 10 0

50

100

150

200 250 300 Aerostat Drag, kg

350

400

450

500

Figure 18: Extra Tether required vs. Aerostat Drag

25

The plot suggests that the percent extra tether increases by the square of aerostat drag. Therefore, a small reduction in aerostat drag can significantly decrease the total weight of the system. This demonstrates the importance of reducing the effects of blow-by to reduce weight.

Tether Weight and Resistive Loss The parameters for the Standard tether in Table 4: Weight of tether as a function of the type of reinforcing fiberare assumed once again to illustrate resistive losses in the estimated tether. Specifically, the conductive wire of the tether was assumed to be 33.1% of the total weight of 102.3 kg/km. Two conductive wires were compared: copper and aluminum 1100. For each wire material, contours of resistive loss were plotted with respect to wire diameter and voltage:

Required Wire Diameter for Standard Tether

Required Wire Diameter for Standard Tether

Figure 19: Contours of % Resistive Loss for Copper (left); Contours of % Resistive Loss for Aluminum 1100 (right)

The contour plots suggest that aluminum 1100 is the more efficient, light conductive wire for the study purposes. Aluminum 1100 actually creates more losses than copper at a given wire diameter and voltage. For instance, for a wire diameter of 1 cm and a voltage of 500 volts, the loss is 10% in aluminum 1100 compared to 6% in copper. The low density of aluminum 1100 allows it to be almost twice the diameter of copper, which results in lower losses at a given voltage. However after careful consideration, due to the elasticity of strength members like Dyneema SK75 copper wires seem more suitable, supporting the Magenntype tether. The tether would produce a loss of about 35%, which is not acceptable. Therefore, the turbine for the DAWT and the Magenn aerostats should be equipped with a generator which outputs a voltage higher than that for ground turbines. This is because heavy gear boxes are needed to increase rotational speed, or heavy generators are needed to increase torque. An advantage of the DAWT is that the velocity increase at the turbine blades increases the tip-speed-ratio, meaning the rotational speed is greater. Another generator option is to use a direct-drive system, which eliminates the need for a gearbox. Eliminating the friction in the gears significantly increases efficiency and decreases wind cut-in speed.

26

Effects of Altitude Variation Altitude of the deployment site is an important variable for any aerostatic design because it affects lift. The lift capacity of the envelope gas decreases with altitude. Therefore, if the same size aerostat is used at higher altitudes than it is intended for, the lift may not be enough to properly support the tether and adverse blow-by will result from this. If the lift is just high enough to maintain height, gas loss due to leakage may drop the lift below the acceptable threshold in duration of time less than the time between maintenance. Both results are undesirable and should not be seen within the design range of operational altitudes. To determine how lift capacity varies with altitude, air properties at altitude are assumed at International Standard Atmosphere (ISA) conditions. For a given aerostat size and weight, the excess lift is a function of the lift capacity, and it depends on altitude. Confluence Lines Most aerostats distribute the tether load across several points along the envelope through wires called confluence lines. Confluence lines main purpose is to avoid excessive load on the envelope membrane at a single point. The confluence lines are then joined with the tether at the confluence point through a pivot which allows the aerostats to rotate freely and align with the direction of the wind. Typically, confluence lines contribute about an additional 1% of envelope mass to the structure (6). Accordingly, materials for confluence lines were investigated and findings are supplied in Table 6. Table 6: Materials for Confluence Lines Material Nylon 6 (39) Cortland plasma 12 rope (24) Density (kg/m^3) 1140.00 980.00 Break Strength 39400 kN/m^2 200 kN Thickness (mm) 0.18 15

Notice plasma rope is introduced here because it defined as the strongest rope per unit weight. Thus, continuing a light weight philosophy, Cortland plasma rope was selected over Nylon 6. Key traits include a polyurethane finish that provides weather resistance and a torque free braided construction. Additionally, Cortland plasma 12 rope is available for a wide range of thicknesses from 1 to 96 mm. Drag due to Confluence Lines To calculate the influence of a confluence line to the drag of an aerostat system, each confluence line (wire) is approximated as a circular cylinder with radius R. Accordingly, the drag on a circular cylinder is approximated as: ( ) where is the coefficient of drag, A the cross-sectional area, q the dynamic pressure, and L is the length of the wire. Thus, it is necessary to further approximate in order to find drag. Subsequently, once the Reynolds number (Re) for free-stream conditions ( and ) is approximated by:

27

( ) in which D is the cylinder diameter, then the drag coefficient for a smooth cylinder as a function of Reynolds number is:

Figure 20: Coefficient of drag as a function of Reynolds number (40); cylinder curve follows closely to circular disk As a comparative analysis, three specific diameter wires were analyzed to determine its drag sensitivity. The wires length was then modeled based upon the systems greatest vertical diameter, in which the wires terminate approximately 1 vertical diameter below the center of gravity (41). Thus, for a general full-scale system confluence wires were approximated using a length of 20 m. Table 7 shows the importance diameter has on the drag of each wire.

Table 7: Drag force produced by confluence lines @ (8 m/s)


Diameter (mm) 2.5 11.0 20.0 Sectional Drag (N/m) 1.115 1.021 1.005 0.101 0.406 0.726 Drag w/ 20 m (N) 2.014 8.115 14.523

Reynolds 1305 5700 10400

Notice that as the Reynolds number varies, the coefficients of drag at these operating conditions are approximately stable. Thus, Table 7 demonstrates the importance of comparing the sectional drag instead of the drag coefficient. Accordingly, it was found that the wires contribute approximately 1.5 % to the total system drag. For example if each confluence wire was 20m long, a total drag force of 64.92 N would be added to the system. Now suppose the free-stream velocity increased to 20 m/s, which typically occurs over the Great Plains. Drag would play a much more dramatic role as shown in Table 8.
28

Table 8: Drag force produced by confluence lines @ (20 m/s)


Diameter (mm) 2.5 11.0 20.0 Sectional Drag (N/m) 1.05 1.00 1.00 0.593 2.484 4.516 Drag w/ 20 m (N) 11.855 49.676 90.320

Reynolds 3200 14000 26000

Number of Lines Required A study was then performed to determine the amount of separate wires that are needed and their effect on drag of the system. For stability purposes the study involves a system comprising of one or more confluence lines. Unfortunately, it seems that there are multiple problems to consider while solving a confluence line balance. Key issues include: Length to Drag ratio of each wire Wires must secure tangentially to surface Secure points affect stability Stress Limit of securing patches Accordingly the first task became finding the number of lines needed. This was accomplished by designing the securing patches and determining their stress limit. To expedite the process, the systems will use an attachment plate design provided by Analysis and Design of Robust Helium Aerostats with a slight modification change to reduce the amount of material exhausted (24).

Figure 21: Attachment Plate (24) Each wire could then be modeled as an 11 mm diameter circular cylinder, the appropriate diameter for an aerostat of this size (24). Moreover, typical system requirements of this size suggest a minimum of 8 total wires placed symmetrically around the aerostat. The attachment plates were simulated as rectangular areas of the envelope in which the thickness increases by 25 mm, in which complete adhesion is assumed.
29

By consequence the addition of the 11 mm wires would then contribute approximately 10 % to the total system drag. Of course, if drag was the main determining factor a smaller wire would be chosen to reduce the contribution to less than 10 %, but in the end the 11 mm wire was chosen to meet stress constraints. Lastly, the length of the confluence wires should be increased only to provide the necessary stability requirements to each aerostat system.

COST METHODOLOGY Coal power plants, despite the millions of tons of CO2 and other harmful gases which are placed into the air, still make up 44.9% of our electricity generation (1). This is not only due to the fact that coal provides millions of jobs, but no energy technology currently available can beat its price of electricity produced. Currently, coal costs $0.04 per kWh, although this does not include the environmental costs associated with it. Average wind power on the ground costs are around $0.08 per kWh (42). Although wind energy has the potential to be cheaper by comparison once installed, the high upfront capital cost leading investors to shy away. The amount of land needed, and the time taken to get a permit, presents more problems. If external costs are included, wind power becomes very competitive, but negative externalities are not being paid for with something such as a carbon tax. The other problems associated with wind are that its simply not consistent on the ground. Until such a tax is imposed, innovations in wind technology can be beneficial. To optimize cost turbine efficiency must be improved, material and installation costs must be reduced, or operating wind speeds must be increased. The latter of the three proves to be an avenue worth pursuing, as winds are much faster and far more consistent at higher altitudes, and wind velocity is the most important driver for performance. Solar-energy generation is expensive due to the upfront cost of the silicon cells. Until photovoltaic cells prices go down, solar energy will continue to be a costly alternative. Solar panels are not formfitting on every house, and some roofs dont face the proper direction to get the full efficiency of the sun. Cloud cover can also hinder the solar output, as can buildings, and there is of course the problem with nighttime. With a mobile solar balloon, there is no need for a proper roof to fit the panels, no buildings to block sunlight, and a greater efficiency can be achieved with more area being exposed. The oblate spheroid shape was used in the balloon analysis. The following concepts were analyzed for their cost: 1. The diffuser augmented wind turbine 2. The Magenn MARS concept 3. The oblate spheroid solar balloon The following equation was used to calculate how much each design would cost over its lifetime:

(5.5) where:

LEC = Average lifetime levelized electricity generation cost It = Investment expenditures in the year t Mt = Operations and maintenance expenditures in the year t Ft = Fuel expenditures in the year t Et = Electricity generation in the year t
30

r = Discount rate n = Life of the system

General Manufacturing and Labor Assumptions Producing more of the same product reduces the average direct cost because of improved labor productivity, and mass-purchasing of materials. Analysis provided by David J. Seymour, Ltd. for manufacturing has determined that producing 100 models from the original prototype would reduce costs by 5%. A 30% discount for mass-purchase of materials was also included. The combined effects of labor and material cost reductions would lower the 100th unit to approximately 78% of the prototype. Averaging this out for all 100 units would put the design at 80% of the prototype for those 100 units. No further cost benefit has been assumed for greater than 100 units, but the trend for reduction in costs is expected to continue. (43)

General Winching Cost Assumptions The winching apparatus for the three designs was scaled such that it would approximate at $0.80 per watt of power output. This factor is only accurate up to a model of 100 kW. A key assumption was a price floor of $50,000 regardless of the wattage. Using known data points, as displayed in Figure 22, linear and exponential trend lines were drawn to determine an equation to find a more accurate relation between the wattage and cost of the winching system.
400000 350000 300000 250000 Cost ($)

200000
150000 100000 50000 0 0 200 400 600 800 Power output (kW) 1000 1200

Figure 22: Cost Trend Lines for Winching System With the equations of these lines now known, the Table 9 displays the predicted value, the known value, and the error for each.

31

Table 9: Winching System Cost Trends


Power Output (kW) 30 100 500 1000 30 100 500 1000 Known Cost (Thousands $) 50 80 160 320 50 80 160 320 Predicted Cost (Thousands $) 50.7 69.6 177.4 312.3 61.1 69.3 142.4 350.1

Trend

Percent Error 1.37% 13.0% 10.9% 2.41% 22.2% 13.4% 12.4% 2.72%

Linear

Exponential

By taking the square root of the squared sum of the percent errors, the total percent errors for the linear and exponential approximations are 17.2% and 28.8%, respectively. For the power outputs examined in this study, it is therefore apparent that the linear approximation is more accurate. It is also important to note that the known costs are rough estimates provided by communication with the Magenn concepts designer, Rivard P. The linear approximation is only needed for the scaling of the Magenn design for 250 kW and above 1 MW, and for the 250 kW solar balloons. For each design, the strips were included as part of the overall envelope cost. They were determined to be at 8% of the initial envelope. Electricity Generation Sources for Comparison

Figure 23: Comparative LEC for different types of energy sources

32

SHROUDED TURBINE CONCEPT ANALYSIS Shape Optimization Several cross-sectional shapes were developed for an analysis of shroud envelope geometry. The shroud cross-sections are intended to resemble Altaeros prototype design, while incorporating characteristics of a DAWT cross-section from Widnall, and an annular airfoil cross-section from Fletcher, Figure 24. The outer diameter of the shroud was kept constant or slightly decreasing in the streamwise direction, as in the Altaeros prototype. By not increasing the outer diameter in the streamwise direction, as in Widnalls design, the area perpendicular to the flow is minimized, thus maintaining low form drag. The shapes of the airfoils, specifically the thickness distribution of the airfoil, were matched to the annular airfoils used in a NACA experiment by Fletcher (44) to determine their aerodynamic characteristics. If the shroud airfoils are sufficiently similar, the data from this experiment can presumably be used to characterize the shrouds. Finally, the DAWT from Widnall can be investigated by incorporating the most notable airfoil characteristics: a wedge-shaped trailing edge, a maximum thickness at about 15% chord, and a unique leading edge.
MIT Shape 20 15

10

-5

-10

-15

Figure 24: From left to right: Altaeros prototype (45), NACA annular airfoil (44), and a DAWT shroud from Widnall. (10)
-20 -10 -5 0 5 10 15

20

Shroud shapes are represented by a 2-D airfoil in the x-y plane with a unit chord. The geometry of an airfoil is determined by the location xt and magnitude t of its maximum thickness relative to chord. A 3-D shroud is produced by scaling the airfoil to a desired chord length c, translating the airfoil in the negative y-direction, and revolving the airfoil about the x-axis. Equations were developed to represent four airfoil shapes, labeled as Shroud1, Shroud2, Shroud3, and Shroud4. The four shapes are presented in Figure 25, for xt = 0.3 and t = 0.15. The equations for each shape are given with subscripts describing the location of the curve: up = upper surface, lo = lower surface, le = leading edge, mi = middle surface, tr = trailing edge.

33

Shroud Airfoil Shapes 0.8 0.7

Shroud1
0.6 0.5 0.4

Shroud2

y
0.3

Shroud3
0.2 0.1

Shroud4
0 0 0.1 0.2 0.3 0.4 0.5 x 0.6 0.7 0.8 0.9 1

Figure 25: Shroud airfoil shapes used for analysis; xt = 0.3 and t = 0.15. Shroud1 is a simple, base-line shape which incorporates the types of equations used in the GNVR shape, specifically the elliptical leading edge. The GNVR shape is optimized for low drag and manufacturability, so using similar equations is likely to yield similar benefits. The curves for Shroud1 are given in Equations 6 through 9. * ( ) + ( )

) ]

( )

) ]

(8) ( )

Shroud2 incorporates an altered trailing edge, as compared to Shroud1. A trailing edge wedge like that seen in the Widnall airfoil is introduced to decrease the wedge angle , which has a positive effect on aerodynamic efficiency and stability. Equations 10 and 11 of the new middle and trailing upper surface replace Equation 8 for Shroud2: * ( ) + ( )

34

)]

Hoerner emphasizes the importance of a sufficiently sharp leading edge for efficient lift characteristics (40). At large trailing edge angles, separation can result in a negative lift curve slope at low angles of attack, which would destabilize the aerostat. The equation for the trailing edge angle for Shroud2,Shroud3 and Shroud4 is given below: [ ( )] ( )

Hoerner suggests that a standard airfoil has a wedge angle with a tangent equal to its thickness with respect to chord. For the example airfoil in Figure 26, this case corresponds to . It can be seen that at higher wedge angles, the lift curve slope is reduced, and a switching angle is developed. The switching angle is defined as the non-zero angle of attack above which positive lift is achieved. For the shroud shapes, it can be shown that for . Because the effects of going slightly beyond this point are small, the maximum allowable should be considered about 40% of chord, marked with a dotted line in Figure 26, which is an 11% increase in past the suggested value.

xt = 40 % Figure 26: Effect of wedge angle, , on airfoil lift. Upper left plot: lift curve slope vs. tan(). Lower left plot: switching angle vs. tan(). Right plot: lift curve displaying a negative slope and switching angle. (40)

With regards to stability, it is beneficial in an aerostat to have the center of buoyancy as far forward as possible. By trimming off some of the thickness in the trailing edge, the center of buoyancy is moved forward for a given chord length. Shroud2 is expected to have superior stability characteristics because of this.

35

Shroud3 incorporates an altered leading edge which resembles that of Widnalls airfoil shape. The curve equations from Shroud2 are maintained except for the upper and lower leading edge equations. The new leading edge is given in Equations 13 and 14: [ ( ) ] ( )

) ]

By moving the stagnation point further from the longitudinal center of the shroud, a greater mass of air is theoretically allowed to flow through the shroud because the stagnation point, or ring in this case, acts as a boundary between air flowing through and air flowing around the shroud. This is presumably the intent of Widnall in choosing her airfoil shape; to increase mass flow and thus power output. The altered leading edge could pose problems, however, if flow separation proves to be a significant constraint on design, because the flow would experience a higher adverse pressure gradient through the duct. Additionally, a sharper leading edge might be more susceptible to dimpling because stagnation pressure is distributed over a smaller area. Shroud4 is a symmetrical airfoil intended to resemble standard airfoil shapes like the NACA series symmetric airfoils. The leading edge is described by Equations 6 and 7 also used in Shroud1 and Shroud2, while the middle and trailing surfaces are given by Equations 15 through 18: * ( ) + ( )

( [ [ ( (

) )] ) ]

( ( (

) ) )

The resemblance of Shroud4 to the NACA series airfoils is evident in Figure 27, which shows the shapes expanded in the y-direction for detail. The y-coordinates of the shroud are within 4% of those of the NACA series airfoil. The most significant difference is in the trailing edge, which is trimmed down intentionally to improve stability.

36

Shroud4 Compared to NACA Series 0.2 Shroud4 NACA 0015 0.15

0.1

y
0.05 0 -0.05

0.1

0.2

0.3

0.4

0.5 x

0.6

0.7

0.8

0.9

Figure 27: Matching of symmetric NACA airfoil

A comparison of thickness distributions of the shroud shapes and two standard airfoils is shown in Figure 28. The thickness distributions are identical for Shroud2, Shroud3, and Shroud4. It can be seen that a significant reduction in thickness in the trailing portion of the airfoil is achieved as compared to Shroud1. Because the thickness distributions are similar, it is assumed at this point that the shroud shapes will have similar aerodynamic characteristics of the NACA series and Clark Y airfoil.
Thickness vs. Chord 0.16 0.14 0.12 0.1 0.08 Shroud1 Shroud2,3,4 NACA 0015 Clark Y

t
0.06 0.04 0.02 0 -0.02

0.1

0.2

0.3

0.4

0.5 x

0.6

0.7

0.8

0.9

Figure 28: Thickness distributions of shroud shapes and standard airfoils.

37

Aerodyanamics of Annular Airfoils Experimental data on the aerodynamic characteristics of annular airfoils is available from NACA (44). The airfoil used is a Clark Y airfoil of 11.7% thickness as seen in Figure 28, which shows the system of axes used in the experiment. The Reynolds number of the flow was between 0.704*10 6 to 2.11*106. The experiment was conducted on five annular airfoils with various aspect ratios, denoted by , which is the ratio of the inner diameter of the duct to the chord length of the duct airfoil. The aspect ratios tested which are relevant to the inflatable shroud analysis are 1/3, 2/3, 1, and 1.5. The results for the aerodynamic center , lift curve slope and zero-lift moment coefficient were collected in Matlab. A relationship for each characteristic as a function of aspect ratio was developed by curve fitting the data points. The resulting curve fits are shown in Figure 29, Figure 30, Figure 31
Aerodynamic Center vs. Aspect Ratio 0.4

0.3

0.2

0.1

-0.1

-0.2

0.5

1.5 d/c

2.5

Figure 29: Aerodynamic center locations at various aspect ratios from NACA experiment Equation 19 was determined from the data points in Figure 29 using a rational model fit with a coefficient of determination R2 of 0.9995: ( ) ( ) ( ) ( )

The aerodynamic center is seen to increase with aspect ratio. At low aspect ratios, below about 0.4, the aerodynamic center is located in front of the leading edge, similar to a slender body without fins. For example, Hoerner estimates that a slender airship hull without fins would have an aerodynamic center about one hull length in front of the nose (40). The instability of such slender envelope shapes dictates the need for stabilizing fins. Thus, if the aspect ratio of an annular airfoil is too low, fins would also be required. Because the inner diameter is constrained by the turbine diameter, the chord length is inversely proportional to the aspect ratio. To reduce the need for fins, its predicted that the thickest shroud, and therefore the shortest chord length, possible will be desired.

38

Lift Curve Slope vs. Aspect Ratio 0.1 0.09 0.08 0.07 0.06
L

0.05 0.04 0.03 0.02 0.01 0

0.2

0.4

0.6

0.8

1 d/c

1.2

1.4

1.6

1.8

Figure 30: Lift curve slope at various aspect ratios from NACA experiment Equation 20 was determined from the data points in Figure 30 using a second-degree polynomial fit with a coefficient of determination R2 of 1.0000: ( ) where is in units of deg-1 ( ) ( )

The lift curve slope is seen to increase with aspect ratio, as with plane rectangular wings. Fletcher found the lift curve slope to be twice the lift curve slope of a plane rectangular wing of the same aspect ratio. Also, he found that for aspect ratios below 2.4, annular airfoils have larger maximum liftdrag ratios than plane rectangular airfoils. For the shroud design, aspect ratios below 1 are expected, so this finding is encouraging to the feasibility of the shroud as an efficient lifting body. Fletcher also concludes that annular airfoils are less stable in stall than plane rectangular airfoils. Stall is not intended to ever occur with an aerostat, which operates at small angles of attack, so this disadvantage should not affect the shroud design.

39

Zero Lift Moment vs. Aspect Ratio 0.03 0.025 0.02 0.015 0.01
MO

0.005 0 -0.005 -0.01 -0.015 -0.02

0.2

0.4

0.6

0.8

1 d/c

1.2

1.4

1.6

1.8

Figure 31: Zero-lift moment coefficient at various aspect ratios from NACA experiment The zero-lift moment coefficients were not provided directly, so they were calculated from the moment coefficient at quarter-chord, , the lift coefficient, , and the aerodynamic center location, , using a simple moment transfer equation: ( ) ( )

Equation 22 was determined from the data points in Figure 31 using a rational model fit with a coefficient of determination R2 of 0.9999: ( ) ( ) ( )

Because the Clark Y airfoil used to obtain the results is similar in shape to the shroud airfoils, it is assumed that these equations are valid for shroud shapes with aspect ratios between 1/3 and 1.5. A large assumption is that the effect of a diverging duct seen in the shroud shapes does not influence the results. Fletcher also provides an equation for the drag of an annular airfoil: ( )

( ) The slope of the drag curve, , is needed for stability calculations:

( ) The zero-lift drag, surface:

, can be estimated using a formula suggested by Hoerner (40) for a lifting

40

where is the friction coefficient, is the wetted surface area, is the reference area, and is the form factor of the airfoil. The reference area is the minimum inner diameter multiplied by the chord length: . (46). The skin friction is estimated with by the Schlicting formula for turbulent flow over a flat plate: ( ) ( )

where the Reynolds number, Re, is a function of the shroud chord length. The form factor, FF, of the airfoil is found using the following equation from Hoerner (40): ( )

Shroud Geometric Parameters The surface area of a shroud is calculated using the equation for a surface of revolution (46):

where SA is the surface area of the shroud envelope, which extends longitudinally from x = 0 to x = c. This solution is found by integrating differential elements, or slices, about the longitudinal x-direction. If these slices are multiplied by a moment arm in the x-direction and integrated in a similar way, the first moment of area about the y or z axis is found. Dividing the moment of area by the surface area gives the location of the center of area, or if envelope density is constant, the center of gravity of the envelope. Thus, the equation for the longitudinal center of gravity, , of a uniform density envelope is given:

The calculation of the overall center of gravity, shroud envelope weight, , the turbine weight, ( (

, takes into account three weight components: the , and the lifting gas weight, . ) ) ( )

where the turbine is considered as a point force at the location of maximum thickness, , and the center of gravity of the lifting gas, assuming uniform density, is also the center of buoyancy, . The volume of the shroud is calculated using the equation for a volume of revolution (47):

41

where V is the volume of the shroud envelope, which extends longitudinally from x = 0 to x = c, and and are the upper and lower coordinates of the shroud airfoil . This solution is found similarly to the surface area; by integrating differential disks about the longitudinal x-direction. By multiplying each disk by a moment arm and integrating, the first volumetric moment is found. Dividing by the volume yields the location of the center of buoyancy, : | | ( )

These equations were solved in Matlab using symbolic integration where possible. However, the symbolic integration in Equations 28 and 29 does not have an explicit solution for the elliptic and parabolic curves. In these cases, a vector fit was performed using a piecewise cubic Hermite interpolating polynomial fitting method. To make the fit equation accurate, vector points were more concentrated at the leading edge of the shroud where the slope of the curves is steepest. To demonstrate the accuracy of these methods, the calculations were performed on a simple shroud shape with established parameters: a torus. To create the torus shape, the equations of a circle in the x-y plane were developed: The resulting 3-D shape by revolving the circle around the x-axis is displayed in Figure 32: ( )

Figure 32: Torus shape example The formulas for the volume and surface area of the torus are given in Equations 34 and 35: ( ( ) )

Calculations were performed for the case where R = 1 and r = 0.5. The exact solutions, calculated with Equations 34 and 35 are SA = 19.7392088 and V = 4.9348022. The volume was then found with Equation 31 using symbolic integration. The volume was found with a relative error magnitude of 10 -12. The surface area was found using a vector fit with Equation 28, using a vector of 10,000 points. If the points are evenly distributed along the x-axis, the relative error magnitude is 10-4. When the points are concentrated towards the edges of the torus, the relative error magnitude is reduced to 10 -8. Equations 29 and 32 yield and , as expected, with negligible error. The results of the torus

42

calculations demonstrate the accuracy of the shroud calculations and confirm that the shroud characteristics can be found without significant error. To develop a wide variety of shroud geometries for analysis, the parameters xt and t were varied for each shroud. For a given xt and t, the necessary volume and surface area were calculated to provide the aerostat platform 25% excess lift. The assumptions made in sizing the shrouds are as follows: Envelope material properties are for mid-grade Vectran, as listed in Material Properties section. 20% extra weight is added to the envelope weight to account for elastic strips, seams and attachments. Tether properties are for the Standard tether, as listed in Tether section. Tether length is 30% greater than design height of 500-m to account for blow-by. Deployment site is at mean sea level. Internal pressure is equal to ambient pressure. Lifting gas is helium. Payload weight is equal to the turbine weight (2750 kg). There is a small, 10-cm gap between the turbine blades and the shroud.

The results for thicknesses of 5% to 30% ,at increments of 5%, at four different maximum thickness locations, 10%, 20%, 30% and 40%, were compiled into an Excel spreadsheet for each shroud shape. The results can be plotted to note the trends in the data. To start, the shroud shapes were compared by plotting various characteristics versus thickness at a fixed xt of 30%:
Surface Area - Volume Ratio vs. Thickness for x t = 0.3 1.4 1.3 1.2 1.1 Shroud1 Shroud2 Shroud3 Shroud4

SA/V

1 0.9 0.8 0.7

0.05

0.1

0.15 t

0.2

0.25

0.3

Figure 33: Surface area volume ratio vs. thickness compared for each shroud shape

Figure 33 shows that for all shroud shapes, the surface area to volume ratio is reduced by about half as airfoil thickness is increased from 5 to 30 percent chord. Minimizing this ratio reduces the relative weight of the envelope and reduces gas losses. Thus, large thicknesses are favorable. Comparing the shroud shapes, Shroud1 and Shroud4 have similar surface area to volume ratios which are lower than the other two shrouds. Therefore, Shroud1 and Shroud4 have the best envelope weight and gas retention properties.
43

Center of Buoyancy vs. Thickness for x t = 0.3 0.45 0.445 0.44 0.435
b

0.43 0.425 0.42 0.415 0.41 0.05

Shroud1 Shroud2 Shroud3 Shroud4

0.1

0.15 t

0.2

0.25

0.3

Figure 34: Center of buoyancy vs. thickness compared for each shroud shape The center of buoyancy can be seen to vary little with thickness for each shroud shape. As predicted, Shroud1 has a center of buoyancy furthest aft because of its thicker trailing edge. Interestingly, it can be seen, by comparing Shroud2 and Shroud3, that by altering the leading edge of the airfoil, xb is slightly reduced. Shroud4 yields the most favorable centers of buoyancy, but the difference is relatively small (about 0.5% chord).
Center of Gravity vs. Thickness for x t = 0.3 0.375

0.37

0.365

Shroud1 Shroud2 Shroud3 Shroud4

0.36

x
0.355 0.35 0.345 0.34 0.05

0.1

0.15 t

0.2

0.25

0.3

Figure 35: Center of gravity vs. thickness compared for each shroud shape The center of gravity is shown to move forward, with increasing thickness. This is due to a lighter envelope as thickness increases which gives the turbine more influence over the center of gravity location. There is not much variation between the shroud shapes (less than 1% chord).

44

Throat-Exit Area Ratio vs. Thickness for x t = 0.3 2.8 2.6 2.4 2.2 Shroud1 Shroud2 Shroud3 Shroud4

Area Ratio

2 1.8 1.6 1.4

0.05

0.1

0.15 t

0.2

0.25

0.3

Figure 36: Throat exit area ratio vs. thickness compared for each shroud shape

The throat to exit area ratio is a measure of the power amplification expected at the turbine. Figure 36 shows that area ratios for Shroud4 are about 2/3 of those for the other three shroud shapes. Therefore, less power amplification is expected with Shroud4. However, many studies on DAWTs have shown the difficulties in controlling flow separation with large area ratios. It remains to be determined whether larger area ratios will be a benefit or a disadvantage for this reason. The results were then compared for a single shroud shape, Shroud3, to see the effects of varying xt. The plots of chord length, center of gravity, and center of buoyancy are as shown:
Chord Length vs. Thickness for Shroud3 55 50 45 40 x t = 0.1 x t = 0.2 x t = 0.3 x t = 0.4

c (m)

35 30 25 20 15 0.05

0.1

0.15 t

0.2

0.25

0.3

Figure 37: Chord length vs. thickness compared for several max thickness locations

Figure 37 shows how the chord length of a shroud varies with t and xt. Chord length is slightly reduced by increasing xt, but is more heavily inversely related to t. As thickness is increased from 5 to 30 percent of chord, length is reduced by a factor of about 3.
45

Center of Gravity vs. Thickness for Shroud3 0.45 x t = 0.1 0.4 x t = 0.2 x t = 0.3 x t = 0.4 0.35

x
0.3 0.25 0.2 0.05

0.1

0.15 t

0.2

0.25

0.3

Figure 38: Center of gravity vs. thickness compared for several max thickness locations

Figure 38 shows how the center of gravity is heavily influenced by xt because the turbine weight is shifted. The center of gravity is always behind the turbine with the distance behind the turbine increasing with xt.
Center of Buoyancy vs. Thickness for Shroud3 0.46 0.45 0.44 0.43 0.42 x t = 0.1 x t = 0.2 x t = 0.3 x t = 0.4

x
0.41 0.4 0.39 0.38 0.37 0.05

0.1

0.15 t

0.2

0.25

0.3

Figure 39: Center of buoyancy vs. thickness compared for several max thickness locations In Figure 39, the center of buoyancy is shown to move forward an increment of about 2% chord for every 10% chord that the maximum thickness is moved forward. It appears from these results that the maximum thickness should be as far from the leading edge as possible to move the center of buoyancy and gravity forward. This furthest location is 40% chord due to the constraint of the tailing edge wedge angle.
46

Aerodynamic Analysis To obtain information on the pressure distribution on the surface of the shroud airfoils in a uniform flow, coordinates for Shroud1,2,3 and 4 were used in a vortex panel applet developed at Virginia Tech by Devenport (48). Values of xt =0.4 and t=0.25 were used. The resulting pressure distribution plots for each shape are shown in Figure 40:

Shroud1

Shroud2

Shroud3

Shroud4

Figure 40: Vortex panel pressure distributions for each shroud airfoil at zero angle of attack

Some important aerodynamic characteristics of the three dimensional shrouds can be inferred from these two dimensional pressure distributions. But first, to help bridge the knowledge gap from two to three dimensions without the use of CFD modeling, the 2-D pressure distributions were determined for two previously analyzed shroud shapes: the airfoil used in Widnalls report, which will be called the Widnall airfoil, and the airfoil used in the NACA experiment, the Clark Y airfoil. The vortex panel pressure distribution on the Widnall airfoil, shown in Figure 41, can be used for a direct comparison of 2-D and 3-D.

47

Figure 41: Vortex panel pressure distribution Widnall

Widnall calculates in her report the pressure distribution on the surface of the shroud for a throatto-exit area ratio of 2. Her distribution, displayed in Figure 42, labeled =0, is shaped much like the vortex panel distribution, with a minimum pressure at about 15% chord. The minimum Cp in the 3-D case, however, is about 1.5 times the magnitude of the minimum Cp from the vortex panel results. Therefore, its assumed that for any of the shroud shapes, the shape of the 2-D pressure distribution is maintained in the 3-D case. Furthermore, the Cp values on the inner/suction surface will be amplified by some factor which is a function of the throat-to-exit area ratio.

Figure 42: 3-D pressure distribution Widnall. (10) Also shown in Figure 42 is the effect of momentum extraction from the turbine on the surface pressure distribution. The parameter is a non-dimensional velocity discontinuity at the trailing edge of the shroud, which is a result of a pressure drop at the blade plane (seen as gaps in the pressure distribution curves). This discontinuity creates a trailing cylindrical vortex sheet of strength , which presumably results in a drag penalty induced by the vortex sheet as well as the pressure drop at the turbine. Thus, if drag is to be minimized, low values of should be considered. For values lower than 0.5, a greater corresponds to a larger momentum change, or disc load, at the blade plane. It can be seen in Figure 42 that as increases from zero, the minimum pressure as well as the slope is reduced in magnitude. This means an improvement in the separation characteristics of the shroud, because the separation criterion relate directly to these two values, as seen in Equation 36:

48

( )

[ ( )] [ ( )] where if , then ( ) ; if , then ( ) [ ( )], and finally if [ ( )] , then separation does not occur (49). Therefore, increasing the turbine disc loading is a tradeoff; on one hand, it is coupled with increased power output, and improved separation characteristics, but on the other, it increases the drag on the aerostat and thus blowby. The turbine disc loading must be carefully controlled to balance these tradeoffs. The discontinuities in the pressure distribution curves in Figure 42 are the result of a theoretical actuator disk. The flow past real turbine blades would undergo a continuous pressure drop, beginning before and ending after the blade plane. This is a critical part of the flow in the suction surface of the shroud because the turbine tends to want to suction the flow off the surface of the shroud which can cause separation. Its essential here to maintain an energized flow close to the surface which resists separation. This can be can be achieved by providing a gap between the turbine blades and the shroud surface, or by ensuring a turbulent boundary layer at the blade plane. Foreman and Gilbert (8) suggest that a short constant area section is needed downstream of the turbine to initiate the core flow diffusion process against the adverse pressure gradient. Such methods should be considered if separation at the blade plane becomes an issue. The ClarkY airfoil used in the NACA experiments (11.7% thickness) was also ran through the vortex panel applet. Along with a pressure distribution plot, the applet provides a lift coefficient and a moment coefficient about the quarter-chord at a given angle of attack. These results were used to calculate the lift curve slope, Cl, and the aerodynamic center with reference to chord, xac, of each airfoil. The aerodynamic center was found using a simple moment transfer equation from the quarter-chord. Also recorded was the surface velocity at the location of maximum thickness, ( ) which gives an indication of how much power amplification the shroud will provide to the turbine. The results are shown in Table 10. Table 10: Comparison of airfoil characteristics Shroud1 1.42 xac Cl 0.239 0.123 Shroud2 1.43 0.237 0.122 Shroud3 1.43 0.237 0.122 Shroud4 1.18 0.237 0.122 ClarkY _ 0.235 0.121 Widnall 1.97 0.217 0.128

The results for xac and Cl confirm the assumption that the NACA experimental data on the ClarkY airfoil is valid for the shroud shapes, because the differences are only fractions of a percent. The Widnall airfoil, however, would not have the same behavior as an annular airfoil because its novel shape moves the aerodynamic center forward, and increases the lift curve slope. The surface velocities at maximum thickness,( ) , reveal that the Widnall airfoil creates the largest velocity increase at the blade plane by far. Shroud1, Shroud2, and Shroud3 all induce about an equal velocity increase, while Shroud4 sees slightly less than half the increase (from a free-stream value of 1) of the other three. The power output of a shrouded turbine using the Shroud4 airfoil will be consequently lower. Note that this result for the ClarkY airfoil is not considered important because it is never considered as a diverging duct. The pressure distribution output for the ClarkY airfoil is shown in Figure 43:
49

Figure 43: Vortex panel pressure distribution ClarkY

The resulting pressure distribution is most similar to Shroud3, evidently because they feature similar asymmetrical leading edges. Comparing these distributions with Shroud2, the asymmetrical edge can be seen to eliminate a sharp pressure peak at the leading edge. The same effect is achieved with the Widnall airfoil as well. This pressure minimum may be undesired because it could lead to laminar separation near the leading edge. On the other hand, Hoerner states that under certain conditions, the pressure peak could advantageously make the boundary layer turbulent to delay separation (40). Although a turbulent boundary layer is likely necessary to prevent separation at the trailing edge, this could be achieved in a more controlled way by tripping the flow with turbulators. This way, laminar flow could be maintained on some portion of the suction surface before transition creates high skin friction drag. Therefore, Shroud3 can be assumed to have superior aerodynamic characteristics to Shroud2, as long as turbulators are used when necessary. Comparing the pressure distributions for Shroud1 and Shroud2, the effect of altering the trailing edge can be noted. The larger wedge angle of Shroud1 results in a steeper curve towards the trailing edge. Because Equation 36 for separation criterion is multiplied by an x term, it is better for separation characteristics for steeper portions of the curve to be closer to the leading edge. Therefore, Shroud2 can be assumed to have superior separation characteristics to Shroud1. From this aerodynamic analysis, it is decided that Shroud3 and Shroud4 should be considered the best airfoil shapes. Although Shroud3 has more potential to increase velocity at the blade plane, Shroud4 is a safe bet because it is least likely to experience flow separation with less extreme C p values. Thus, Shroud3 and Shroud4 are specialized for different purposes; Shroud3 is an efficient power augmenter and Shroud4 is an efficient lifting body. Further analysis is required to determine which performs best in a real flow environment.

Stability Analysis of Annular Shroud The static stability of the aerostat can be analyzed by considering the moments and forces acting on the aerostat in a steady free-stream, , at an angle of attack, . These are represented in the following diagram:

50

Bf U O B G W

L M A
aL

Tv T Figure 44: Stability diagram of annular shroud aerostat C (xc,zc) T


h

cx = xg - xc cz = z g - z c

The diagram defines the positions, relative to chord, of the center of buoyancy, , center of gravity, , and aerodynamic center , along with their respective forces and moments. The confluence point is located at (xc,zc) at distances cx and cz from the center of gravity. The center of buoyancy, center of gravity, and aerodynamic center are shown in their desired relative positions for stability. Therefore, the primary goals in improving static longitudinal stability are to reduce , and increase while maintaining an acceptable . A quantitative measure of static longitudinal stability is the slope of the moment coefficient as a function of the angle of attack, , or static stability margin. At an equilibrium state, the moment coefficient of the aerostat, , is zero with a negative such that any small change in angle of attack would produce a stabilizing moment. Rajani, Pant & Sudhakar (6) suggest that a static stability margin of is sufficient to achieve static and dynamic stability. Thus, this is the ceiling value for all stability calculations (i.e. any stability margin equal to or less than this value is considered acceptable). The equations for static stability analysis of an aerostat are provided by Rajani, Pant & Sudhakar (6). Referring to Figure 44, the forces are balanced in the x-z plane to yield the following equations: ( ( where, ) )

51

[ [ ]

( ( (

) ) )

The lift curve is approximated by a sine curve to simulate a loss of lift at high angles of attack: ( where, at low angles of attack , in units of radians, as approximated by Equation 20. , are approximated with Equations , where is the minimum inner )

The coefficient of drag, , and the zero-lift moment coefficient, 22 and 23. All coefficients are with reference to the area diameter of the shroud, and is the chord length of the shroud.

By balancing moments about the confluence point and simplifying, the moment coefficient is given: * [ +( ( ) ) [ ] [ ] ] ( )

A confluence point location is chosen such that the equilibrium angle of attack is independent of wind speed, . The solution for this location is given in Equations 44 and 45. ( ( ) [( ) *( ) )
(

(
)

) ] +

(45)

Finally, the margin of static stability is shown to be [ [ [ ] [ ]( ] ( ) ( ) ) ]

The margin of static stability was calculated for each shroud variation in the Excel spreadsheet of shroud geometries (t varied from 5% to 30%, xt varied from 10% to 40%). A free-stream velocity of = 8 m/s was considered as an expected average wind speed at the aerostat height. The angle of attack was set at = 2, which is a typical trim angle of aerostats. Air properties were used for ISA conditions at 500 meters altitude. All shroud variations were found to have unstable positive stability margins. Therefore, there is absolutely a need for stabilizing fins on the annular aerostat. Stability was found to improve by decreasing t and by increasing xt. Thus, the most stable geometry was found to be t = 0.05 and xt = 0.4. The stability margin at a given t and xt was found to vary consistently between the shroud shapes. The order of stability of the shroud shapes, starting with the most stable is as follows: Shroud4, Shroud3,
52

Shroud2, and Shroud1. At the most stable geometry (t = 0.05 and xt = 0.4): Shroud4 was found to have a margin of stability of = 0.576; for Shroud3, = 0.584. Only these two shroud shapes will be considered for further analysis because they show the best stability characteristics. Furthermore, xt = 0.4 seems to optimize stability, so only this thickness location will be considered further. Stability Analysis of Complete Aerostat An additional force component should be added for the thrust produced by the turbine, also called disk loading. The coefficient of thrust, with respect to the blade plane area, is denoted as . Widnalls estimations of turbine thrust coefficients for a DAWT are up to 0.6 with reference to the blade plane area (10). Typical wind turbines operate at a range of thrust coefficients depending on wind speed, as seen in Figure 45. Foreman and Gilbert (50) suggest that for a DAWT, peak power augmentation is achieved at a disk loading of 0.42. For this analysis, a thrust coefficient of 0.42 with reference to blade plane area is considered.

Figure 45: Thrust coefficients of a variety of HAWT designs vs. wind speed

The thrust coefficient must be corrected to be in terms of the shroud reference area by multiplying it by a ratio of the reference areas as follows: ( ) The thrust is considered to act in the longitudinal direction independent of angle of attack. Thus, for positive angles of attack, the turbine contributes a negative lift component, , and a positive drag component, . A term must also be added to the drag curve slope: . The effect of adding fins to the aerostat platform on stability was estimated by introducing a fin lift curve component, , and adjusting the aerodynamic center, . Its assumed that all other shroud characteristics are unchanged (i.e. the drag of the fins is negligible, as well as effects on buoyancy and center of gravity). For thin fins, the zero-lift drag is approximately the skin friction drag on a thin plate, which is negligible compared to the skin friction drag and pressure drag on the shroud. Also, at small angles of attack, the induced drag of the fins should be negligible. Therefore, as long as the fins are sufficiently thin and lightweight so as to not affect buoyancy and gravity, the effect of their lift is dominant over other forces and this estimation is sufficient. It should be noted that the error in fin size

53

estimation should be an overestimation because moving the center of gravity backwards should have a stabilizing effect. The lift curve slope of the fins can be estimated by an equation from Hoerner (40): ( ( ) )

where ARf is the fin aspect ratio, is the Prandtl-Glauert compressibility factor, and is the sweep of the fins at half-chord. It is assumed that for normal aerostat operation, compressibility effects are negligible and . This lift curve slope must be corrected to find the change in the total lift curve slope with respect to the shroud reference area: ( ). For aerostats, a relatively low fin aspect ratio is desired to reduce bending moments at the root. For this analysis, the fin aspect ratio and sweep are fixed at values which are reasonable for aerostat fins by approximating the shape of fins on a common Lockheed Martin aerostat. The 2-D shape and relative dimensions of each fin is shown in Figure 46.

ct cr = ct = b
1/2

ARf = 0.75 tan 1/2= 0.5

cr Figure 46: Lockheed Martin NE&SS-Akron 420K aerostat fin shape approximation The fin configuration considered is four fins, two horizontal and two vertical, each spaced by 90 degrees about the longitudinal axis. The fins are considered to be mounted behind the shroud on a spar which extends along the longitudinal center. The length of this central spar is fixed such that the aerodynamic center of the fins is two shroud chord lengths behind the shroud leading edge. The stability diagram for this new configuration, including the turbine forces is shown in Figure 47:

54

1/2

, ,

Bf O B T G W

L Tt N M
nL

Tv T C (xc,zc) T
h

cx = xg - xc cz = z g - z c

Figure 47: Stability diagram of annular shroud aerostat with turbine and fins The overall aerodynamic center of the aerostat is now called the neutral point, located at with respect to chord. The location of the aerodynamic center of the fins is , which is set at 2 chord lengths. The neutral point can be calculated by Equation 48: ( ( ) ) ( )

Fin sizing was performed on Shroud3 and Shroud4 for = 0.4 and a thrust coefficient of 0.42. A fin size was chosen by varying the span of the fins, , until a margin of stability below the ceiling value of -0.246 was achieved. The necessary span was found to the nearest 0.1m. The results were then plotted as functions of thickness:

55

Fin Span vs. Thickness for x t = 0.4 23 22 21 20 Shroud3 Shroud4

b (m)

19 18 17 16 15 0.05

0.1

0.15 t

0.2

0.25

0.3

Figure 48: Required fin span vs. thickness for stability The results in Figure 48 indicate that smaller fin spans are required with larger thicknesses. Therefore, although lower thicknesses are more stable without fins, larger thicknesses are more easily stabilized with the use of fins.
4 x 10
4

Fin Volume vs. Thickness for x t = 0.4

3.5

Shroud3 Shroud4

V (m3)

2.5

1.5

0.5 0.05

0.1

0.15 t

0.2

0.25

0.3

Figure 49: Required fin volume vs. thickness for stability

The fin volume, , is defined as the distance between the fin aerodynamic center and the shroud aerodynamic center, ( ), multiplied by fin reference area, . This is a common way to demonstrate the effectiveness of tail fins in aircraft. In this case, a minimum fin volume is desired to minimize internal moments on the aerostat due to fins, and reduce structural weight. It can be seen that fin volume is unacceptably high at low thicknesses, and is reduced by a factor of about 6.7 as thickness is increased from 5% to 30%.
56

Neutral Center vs. Thickness for x t = 0.4 0.63 Shroud3 Shroud4 0.62

0.61

x (m)

0.6

0.59

0.58

0.05

0.1

0.15 t

0.2

0.25

0.3

Figure 50: Required neutral center vs. thickness for stability The required neutral center of the aerostat was found to be located around 60% of chord for each shroud. The neutral center of Shroud4 can be located further forward than Shroud3 because of the greater stability of the shroud without fins. The effect of the turbine thrust was found to be substantial in increasing blow by due to drag. With a thrust coefficient of 0.42, drag was increased by an average of about 400% from zero thrust. This results in an increase in the angle of the tether at the confluence point with respect to the vertical from about 15 to 55. If the tether is approximated as a straight line, this corresponds to an increase in extra tether length due to blowby from 5% to 75%. This high extra tether length may be unacceptable. Therefore, the thrust coefficient may need to be lowered to reduce the drag on the aerostat. Adjusting the thrust coefficient is a tradeoff between power output and drag that should be further investigated. These analyses suggest that the symmetrical airfoil of Shroud4 is an optimal shape for buoyancy, aerodynamics and stability. However, Shroud3 is the optimal shape if a larger throat-exit area ratio, and higher power augmentation, is desired. The aerodynamic effects (i.e. flow separation) of the larger throat-exit area ratio of Shroud3 should be further investigated to determine its feasibility.

Material Stress Analysis In order to select the appropriate material fabric for an envelope design, it is necessary to calculate the maximum stresses imposed on the envelope. Accordingly, because the aerostat must maintain its overall shape a positive pressure is needed inside the envelope. Therefore, adding the following three main loadings helps estimate internal pressure in the aerostat (p) as suggested by Gupta & Malik (51): I. Dynamic pressure loading Acts on the front portion of the aerostat envelope and tries to make a depression on the envelope surface. The internal pressure due to dynamic pressure ( ) loading is related by: * + ( )

II.

where is the density at the respective altitude, U is the wind speed, and the internal pressure is typically designed for a 15% overpressure. Aerodynamic loading
57

The loading is created through the aerodynamic pressure differential from applying stability conditions to the aerostat. Most aerostats are operated at an angle of attack within 2-2.5 degrees thus creating the aerodynamic coefficient of pressure (Cp) in the range of 0.30 to 0.35. The internal pressure due to Cp ( ) loading is related by: ( ) * + III. Hydrostatic pressure loading The loading due to the difference in density at the top and bottom of the aerostat can be calculated at maximum diameter relative to the centerline: ( ) ( )

Since the standard aerostat envelope, some version of a sphere, diameter is vastly larger than the thickness of the actual material; it can be considered as a very thin shell. When the envelope is then subjected to the pressure loading p; a hoop stress is created which is illustrated in Figure 51.

Figure 51: Illustration of hoop stress where , is the circumferential hoop stress. Alternately, the DAWT uses a shroud design that resembles two thin walled circular cylinders. Since the shroud is also subjected to both an internal and external pressure; hoop stresses are produced in both the inner and outer envelope shells. Hence, the hoop stress is calculated in terms of the circumferential unit load as shown: * + ( )

in which is the differential pressure, and d is the external diameter. Therefore, adding the three main loadings helps estimate internal pressure in the aerostat (p). This relation thus gives the load that the material fibers will experience in a circumferential fashion. The hoop stress is then compared with the allowable stress values of the envelope design. A factor of safety of 4 is normally kept in the selection of envelope fabric. This is required to take care of the inaccuracies in the calculated material properties such as diurnal temperature variations or degradation in the envelope material due to handling. Table 11 and Table 12 calculate loading calculations at a 500 m altitude with wind-speeds approximately equal to 8 m/s, 20 m/s, and = 1.129 kg/m3.

Table 11: Material loading @ 8m/s; see drag for Cp calculation


Surface Inner, diameter = 16 m Outer, diameter = 24.1 m Dynamic Loading (Pa) 41.55 41.55 Aerodynamic Loading (Pa) 197.05 37.57 Hydrostatic Loading (Pa) 75.54 113.789 Hoop Stress (Pa) 197.05 37.57

58

Table 12: Material loading @ 20m/s; see drag for Cp calculation


Surface Inner, diameter = 16 m Outer, diameter = 24.1 m Dynamic Loading (Pa) 259.67 259.67 Aerodynamic Loading (Pa) 1231.58 234.83 Hydrostatic Loading (Pa) 75.54 113.789 Hoop Stress (Pa) 1231.58 234.83

Through Table 11 and Table 12, it was found that the inner surface of the DAWT would experience the greatest stress. Therefore, the envelope fabric was then designed around the inner surface loading scenarios.

Diffuser Augmented Wind Turbine Cost General Assumptions As with the solar balloons, the DAWT is designed to be off the grid. Estimating prices for installation are therefore negligible, and the prices that need to be considered were included in the operating and maintenance cost. Standard wind turbines typically operate under the condition of 20% of the investment capital cost for O/M costs. (8) However, as this design is higher in the air and exposed to more extreme weather, it will require more attention and maintenance. The envelope will also require more repairs than a traditional horizontal axis wind turbine, thus increasing these costs further. An additional 10% is assumed, bringing the total O/M cost to be 30% of the initial investment. A discount rate of 5% was applied, with a lifetime of 20 years. Wind turbine blades are operational for at least 20 years, and they are the bulk of the cost for this design. Replacing and fixing the balloon is factored into the O/M costs. Discount rates range between 5% and 10%. A 10% discount rate would make the project too expensive, and was therefore not evaluated. Cost/Pricing Analysis Table 13 contains the reference costs for the DAWT design, including manufacturing and labor costs: Table 13: DAWT System Components Reference Cost
Component Gas ($/m^3) Envelope ($/m^2) Power ($/W) Tether ($/m) Confluence lines ($/m) Winch ($/W) Cost $0.28 $25.00 $2.00 $12.00 $7.50 $0.80

59

Using the general manufacturing assumptions taken for all models with the specific assumptions for the DAWT already described, the capital prices for the turbine, including lifetime helium usage, are specified in Table 14 Table 14: Total Investment Cost (in thousands of dollars)
Helium 2 Envelope 61.3 Turbine 200 Tether 8.4 Winch 80 TOTAL 351.8

The winching and mooring system are priced at $0.80 per watt, as stated previously in the analysis. Unlike the solar balloon concept, only a single tether is used. However, the DAWT is lifted to a higher altitude of 500 meters, resulting in a longer tether and therefore more cost. The cost of the tether is fairly negligible with respect to the design as a whole. The helium costs were especially low, as the volume for the shroud was small, and the deflation rate was very slow. If allowed to fully deflate, it would take approximately 20 years. The turbine and power output was priced at $2 per watt. This includes the savings from manufacturing many units and mass-purchase of materials. The cost per watt was estimated by pricing the cost of large-scale wind farms with their respected outputs, and scaling them down. (52) Cost of Electricity Generation According to the initial estimates, the entire volume and area for the shroud as modeled for the wind turbine varies between 3325-3586.3 m3 and 2346.3-2452.1 m2, respectively. With these models, a wind turbine blade of 16 meters in diameter can be placed inside the shroud, leaving 0.5 meters of circumferential clearance. The following equation can be used to estimate the energy generated by the turbine (42) ( where: HWP = power output = density at the specified altitude = sweep area of the rotor blade = average wind velocity at the specified altitude = maximum power coefficient (theoretical maximum of 0.59) )

The resultant life, yearly and daily electricity generation for each capacity factor are listed in Table 15 assuming a lifecycle of 20 years and a 100 kW power output.

60

Table 15: Life, Yearly and Daily Electricity


Capacity Factor 0.25 0.33 0.40 0.45 0.50 Daily Energy (kW-h) 600 792 960 1,080 1,200 Yearly Energy (MW-h) 219.2 289.3 350.6 394.5 438.3 Lifecycle Energy (MW-h) 4,383 5,786 7,013 7,889 8,766

Taking the yearly electricity generation, the total investment costs and operation and maintenance costs from the previous tables and using the algorithm in Equation (5.5), the total lifetime costs are listed in Table 16 Table 16: Shrouded Turbine Total Lifetime Cost
Capacity Factor 0.25 0.33 0.40 0.45 0.50 LEC ($/MW-h) 149.3 113.1 93.3 83.0 74.7 O/M ($/MW-h) 44.8 33.9 28.0 24.9 22.4 LEC total ($/MW-h) 194.1 147.1 121.3 107.9 97.1

Conclusions As seen in Table 16, the capacity factor has a profound effect on the LEC for this design. If the rate is increased from 45% to 50%, the costs decrease by a factor of 10%. This change is observed even more fundamentally if the starting capacity factor is relatively low. While its an important aspect of any energy collection design scheme, the purpose of the higher altitudes is to increase the capacity factor. Higher altitudes not only give higher wind speeds, which would allow for a smaller blade diameter and reduce the overall cost, but the wind speeds are more consistent. Wind energy generation is very potent when the wind is blowing at consistent rates, but its historically unreliable to handle peak hours in certain geographic locations. The higher altitude, and presumably the higher capacity factor, could solve this problem. Equation (53) shows that power output is most affected by the wind speed, as its raised to the third power. It is for this reason that the wind velocity is considered the most important design driver. An increase of 20% in average wind speed can decrease the needed sweep area by as much as 45% for the same power output. This has a lasting impact on the materials, but also on the operation and maintenance. As the needed volume of helium for this power output was relatively low, the lifting gas played a minimal role in the cost. Even if the price of helium were to double, the influence of the tether would still be greater. However, as the volume would be expected to increase at a quick rate if the 100 kW model was scaled up, the gas still plays a more important part in driving the design than does the tether. Its unlikely that this model can be scaled into the MW region; therefore the winching system drives the cost

61

more-so than the gas. Doubling the envelope cost increases the LEC by 20% at a capacity factor of 0.5. It can be concluded from this observation that the envelope drives the cost to a large degree. Table 17 displays the order of design drivers for the DAWT based on the cost: Table 17: Shrouded Turbine Concept Cost Drivers
Rank 1 2 3 4 5 6 Component Capacity factor Wind speed/velocity Envelope Winching System Lifting gas Tether and confluence lines

MAGENN MARS CONCEPT ANALYSIS The Magenn Air Rotor System (MARS) is a horizontal axis Savonius style wind turbine with three vanes and two toroidal endplates. Thus far, Magenn Power, Inc. has produced a small scale, 30kW prototype that was used as a demonstrator in order to show the viability of the design. However, for the purposes of this study, the design must be evaluated at scales that are closer to the utility level. In order to perform such a comparison, a scaling method must be developed for the MARS in order to boost it to higher outputs and dimensions. Scaling Methodology Specifications for a 30kW prototype of the Magenn Air Rotor System (MARS) were obtained from Magenn Power, Inc. in the form of a spreadsheet outlining the detailed geometry and component lists, along with the associated weights (37). The system was scaled up to larger power outputs based on five basic parameters: the diameter of the cylindrical envelope d (in feet), the diameter to the tips of the vanes D (in feet), the spanwise length of the cylinder L (in feet), the overall power rating of the system P (in kilowatts), and the tether length Z (in feet). The overall proportions of the system were kept constant by locking the relationship between the four parameters governing the physical size of the system: d, D, L, and Z. The system was sized in order to allow a maximum value of the buoyancy fraction, defined as: ( ) ( )

where W denotes the overall weight of the system and B denotes the overall buoyant force generated by the system. The buoyancy fraction is essentially a measure of the excess buoyant lift generated by the system. Magenn desires a buoyancy fraction as near as possible to 25% for safety and maintenance reasons, and so each scaled system was matched as nearly as possible to this desired buoyancy fraction, except when size precluded achieving this value. If this value could not be reached, then the buoyancy fraction of the system was maximized. A spreadsheet was written based upon the one obtained from Magenn to perform the scaling of the prototype system. The inputs for this analysis were as follows:
62

1) Desired power output Z, kW 2) Spanwise length of envelope L, feet The other parameters were linked to these inputs by the ratios exhibited in the data given for the original prototype. The desired power output was specified first, and then the length was iterated until a maximum value of the buoyancy fraction was reached.

Scaling Results The original 30kW prototype was scaled up to seven different larger power output ratings: 50kW, 100kW, 250kW, 500kW, 1MW, 5MW and 10MW. Figure 52 summarizes the buoyancy fractions attained at each power rating, where the reference system P and reference buoyancy fraction Fb are the values for the 30kW system model: 30kW and 25% buoyancy fraction. Most of the systems attain around a 25% buoyancy fraction, which was the desired value (37). It is apparent that the much larger systems cannot sustain nearly as high a buoyancy fraction as the smaller systems, which could affect the ability to keep the system aloft for long periods of time. The size of such systems also increases greatly with power rating in order to achieve the desired buoyancy fraction, as shown in Figure 53. The largest (5MW and 10MW) systems do not reach the desired 25% buoyancy fraction, and as such were simply maximized with respect to buoyancy fraction. The implications of this are clearly apparent in Figure 53, with these two largest systems rising to over six-fold of the spanwise length of the 30kW system. This is equivalent of over 500 feet length, compared to a length less than 200 feet for the 1MW scale.

Figure 52: Summary of maximum buoyancy fractions at different power scales (data points shown for 30kW, 50kW, 100kW, 250kW, 500kW, 1MW, 5MW, 10MW, left to right). Larger systems cannot obtain the desired buoyancy fraction of 25%

63

Figure 53: Summary of Spanwise envelope length (data points shown for 30kW, 50kW, 100kW, 250kW,
500kW, 1MW, 5MW, 10MW, left to right)

Aerodynamic Analysis In order to approximate the additional aerodynamic forces created through rotation of the vanes in the MARS, a two dimensional slice of the cross-section is taken as a circular cylinder. Typically, flow over a circular cylinder will follow its surface to approximately half way around the body and then separate causing a large wake. This wake creates a lot of drag therefore if the wakes strength is reduced, then the drag caused by the wake is also reduced. Consequently, consider the circulatory flow about a circle shown in Figure 54.

64

Figure 54: Flow about a circle with two different degrees of circulation, a) moderate circulation, b) circulation so strong that the two stagnation points coincide. (52) In Figure 54 two different levels of circulation are shown; that of moderate circulation and the limiting case in which the circulation is so strong that front and rear stagnation points coincide. If the reference chord is taken as the diameter of the circle, the limiting is easily calculated as: ( )

No large wake is created behind the cylinder, and thus the drag is significantly reduced. The net lift force created by the circulation from the rotation of the envelope and vanes is called Magnus lift, and is one of the major selling points of the MARS system, as it reduces the required gas volume of the system by reducing the dependency of the sytem on buoyant lift. The Magnus lift on a rotating cylinder can be calculated according to the Kutta-Joukowski Theorem by the following: ( ) ( )

according to Bertin and Cummings (53), where R denotes the cylinder radius and denotes the rotational speed of the cylinder in radians per second. We can then calculate the lift coefficient for the cylinder: ( ) ( )

where D denotes the diameter of the cylinder and L denotes the spanwise length of the cylinder. Based on data from the MARS 50x prototype, a rotational speed of about 7RPM can be optimistically estimated; this translates to about 0.73 radians per second (37). Assuming that the aerostat is moored at a 300 meter (1000ft) altitude as per Magenns target altitude with a wind speed of 10m/s (to achieve 100% power

65

output), the Magnus lift force generated by any scale MARS system can be easily estimated. The lift coefficient obtained was 0.0092 for all sized models. The lift to drag ratio for the MARS was also obtained through data from the MARS 50x prototype as around 0.71; this value is less than one, since the MARS is a Savonius style wind turbine, and thus it is a predominantly drag based system. The lift estimations calculated previously could now be used to obtain an estimation of the drag incurred on the system. The results of the aerodynamic estimations on the MARS are shown in Table 18 and Figure 55. They are listed as actual values as well as dimensionless normalized values. The reference values are that of the 30kW system. The reference diameter, Reynolds number, lift per span, lift, and drag are represented as Di, R, l, L, and D, respectively. Table 18: Results of MARS aerodynamic analysis at various power output scales
Power rating (kW) 30 50 100 250 500 1,000 5,000 10,000 Diame ter (ft) 33.1 34.4 37.5 46.1 58.7 80.3 225.8 281.2 Re_D *106 2.06 2.13 2.33 2.86 3.64 4.99 1.40 1.75 Di/Di & R/R 1.00 1.03 1.13 1.39 1.77 2.42 6.80 8.50 l (lb/ft, span) 29.1 31.4 37.3 56.5 91.5 171.6 1355.7 2101.9 Excess lift % 72.6 72.6 72.6 72.6 72.6 72.6 63.7 48.5 L/L & D/D 1.00 1.12 1.45 2.70 5.57 14.30 317.65 613.21

l/l 1.00 1.08 1.28 1.94 3.14 5.90 46.59 72.23

L *103 (lb) 2.40 2.71 3.51 6.53 13.48 34.60 768.71 1484.0

D *103 (lb) 3.41 3.82 4.95 9.20 19.0 48.7 1082.0 2090.0

66

Figure 55: Scaling of the diameter Di, lift per span l, total lift L and drag D

L/W % describes the ratio of the Magnus lift generated by the system to the overall weight of the system, and Excess lift % describes the ratio of the total lift (buoyant and Magnus lift) to the overall weight. Clearly, the excess lift generated by the system, and hence the relative tension in the cable, drops off after the 1MW point is reached. This is an indicator of the possible impracticality of the larger systems; the necessary weight of the power generation system increases faster than the available lift.
Tether Profile and Mooring System The MARS is advertised by Magenn as being a very stable platform in high winds and weather conditions. Part of the design approach to achieving this is to utilize a taut, high tension tether that will allow very little travel of the aerostat in windy conditions. Magenn has specified a tether angle of 32o with the vertical direction for its MARS 50x prototype, and seeks to maintain approximately this value for larger designs (37). A schematic of this tether profile is shown in Figure 56. Using the lift and drag values calculated in the previous section, the tension force T on the cable was calculated as: ( )

where L denotes the lift force and D denotes the drag force on the envelope and the tether and confluence lines. The tether drag is calculated as explained in the section on Drag due to Confluence Lines. It was assumed that the confluence lines accounted for 1/10 of the total tether length, and that each system has two confluence lines, as in the prototype. Based on the methods developed in the Tether/Mooring
67

Section, a mooring system should be obtained that will withstand the cable tension force to a factor of safety of 4; as such, the mooring system resistance weight is taken as 4 times the total tension force on the cable. The results of the tether and mooring analysis for the MARS are shown in Table 19.

Figure 56: MARS Tether Profile (11)

Table 19: MARS tether analysis at various power output scales Power rating 30kW 50kW 100kW 250kW 500kW 1MW 5MW 10MW Tether length (ft) 500 519 566 696 886 1214 3412 4248 Tether drag (lb) 261 271 295 363 462 633 1779 2215 Total drag (lb) 3669.1 4088.3 5244.3 9562.2 19445.0 49371.1 1084464.4 2092316.0 Tension (lb) 4395.3 4905.2 6312.6 11580.0 23659.2 60290.6 1329275.4 2565143.0 Mooring weight (lb) 17581.4 19620.7 25250.5 46320.1 94636.8 241162.6 5317101.7 10260571.8

68

Magenn MARS Concept Cost General Assumptions As with the solar balloons and the DAWT, the Magenn concept is designed to be off the grid. Estimating prices for installation are therefore negligible, and the prices that need to be considered were included in the operating and maintenance cost. Standard wind turbines typically operate under the condition of 20% of the investment capital cost for O/M costs. However, as this design is made almost completely of fabric, it will require more attention and maintenance. The envelope will also require more repairs than a traditional horizontal axis wind turbine, thus increasing these costs further. An additional 10% is assumed, bringing the total O/M cost to be 30% of the initial investment. Using a discount rate of 5%, the Magenn concept is estimated to have a lifetime between 10 and 15 years, and both lifetimes were considered for evaluation. Replacing and fixing the envelope is factored into the O/M costs. Discount rates range between 5% and 10%. A 10% discount rate was ignored, as the other projects could not successfully sustain such a rate. It would therefore be unfair to increase the rate and then evaluate the design on unequal terms and conditions. Cost/Pricing Analysis Table 20 contains the reference costs used for each wind turbine design, including the labor and manufacturing cost: Table 20: Magenn Component Cost
Gas ($/m^3) Envelope ($/m^2) Tether ($/m) Generators (per 25 kW) Storage System ($/W) Structure ($/kW) $0.28 $25.00 $12.00 $15,000.00 $30.00 $1,000.00

Using the general manufacturing assumptions taken for all models with the specific assumptions for the balloon already described, the capital prices for the turbines, including lifetime helium usage, are specified in Table 21. A constant tether price of $8,400 is taken. Table 21: Magenn Total Investment Cost
Power Rating (kW) 30 50 100 250 500 1,000 5,000 10,000 Storage System 3 5 10 25 50 100 500 1,000

Gas 1.8 2.1 2.7 5 10.2 26.4 586 1,131.20

Envelope 37.6 47.1 56.1 84.7 137.4 257.5 2,035.20 3,155.40

Generators 57 95 190 475 950 1,900 9,500 19,000

Winch 50 56 80 110 160 320 1,400 2,750

Structure 10 16.7 33.3 83.3 166.7 333.3 1,666.7 3,333.3

TOTAL 167.9 230.3 337.2 791.5 1,482.8 2,945.7 15,696.3 30,378.4

69

Though the leakage rate for this design could not be estimated, the trends from the solar balloons and the DAWT imply that over the life cycle of the system the gas needed will double from the initial value. For a 50 kW system, $1,050 in helium is needed, and it can reasonably be assumed that this price will double to the indicated value of $2,100 as shown in the table. The volume required for each design was scaled using the provided 30 kW model. With a calculated scaled buoyancy force, it could be divided by the density of the lifting gas (0.164 kg/m3) which resulted in the ending volume. The area was scaled in a similar manner using lengths and diameters. The winch followed the general assumption for the 50 kW, 250 kW, 5,000 kW and 10,000 kW models. All values for estimating costs were provided by communication with the Magenn concepts designer, Rivard P.

Cost of Electricity Generation The resultant life, yearly and daily electricity generation for both 10 and 15 year lifecycles are listed in Table 22 assuming a capacity factor of 0.45. This capacity factor was also provided by Rivard P. Table 22: Magenn Levelized Energy Cost (in thousands of dollars)
Lifecycle (years) Power output (kW) 30 50 100 250 500 1,000 5,000 10,000 30 50 100 250 500 1,000 5,000 10,000 LEC ($/MW-h) 131.4 118.3 104.1 99.0 92.7 92.1 98.2 95.0 97.8 88.0 77.5 73.6 69.0 68.5 73.0 70.7 O/M ($/MW-h) 26.3 23.7 20.8 19.8 18.5 18.4 19.6 19.0 19.6 17.6 15.5 14.7 13.8 13.7 14.6 14.1 LEC total ($/MWh) 157.7 141.9 124.9 118.8 111.3 110.5 117.8 114.0 117.3 105.6 92.9 88.4 82.8 82.2 87.6 84.8

10

15

Conclusions The lifecycle varied on this design with from 10 to 15 years, as specified by Magenn. This differed from the traditional 20 year lifecycle that was analyzed for all of the other designs. This analysis shows that lifecycle is the most important factor to the entire lifecycle cost. A 50% improvement in lifecycle coincides with a 25% reduction in cost. This trend would likely continue if the lifecycle could be extended to the full 20 years. There seems to be a limit at the 5,000 kW power output, but this cannot necessarily be concluded as the 10,000 kW model is also less expensive than the 5,000 design. It would be worthwhile to continue scaling up to see if there is a limit at 1 MW. The models above the 10 MW would not be feasible, but they would provide an answer as to whether there is a limit or not.
70

The capacity factor was kept constant, but like with the DAWT, it can be assumed that if it were increased that it would greatly increase the total energy generated. This is the entire purpose behind putting a wind turbine in the air. The generators affect the cost the most, but their prices cannot really be influenced; the same can be said of the structure and storage systems. Higher efficiency will bring their costs down, and this is the only way to affect that particular cost driver. The winching and envelope costs intersect one another, and if the trend for the winching holds, it is expected that they will intersect once again. However, such a scaling factor is unlikely to be practically feasible, making the envelope a more important cost driver than the winching system for the upper limits. For smaller scale turbines, the winching system carries more weight than the envelope. As the 10 MW was the lowest cost, the ranking system of cost drivers corresponds to that system. Table 23: Magenn Cost drivers
Rank 1 2 3 4 5 6 7 Component Capacity factor Generators/generator efficiency Envelope Structure Storage System Gas Tether and confluence lines

SOLAR AEROSTAT CONCEPT ANALYSIS A solar powered aerostat concept is studied in terms of marketability and feasibility as an energy generating source for comparison purposes. The process includes selecting an adequate envelope shape, the mooring and tether systems after performing an analysis of sizing, aerodynamics, endurance, and cost. The sections that follow show a chronological account of the methods applied and results leading to each design decision that drives the final solar aerostat configuration. Preliminary Analysis Methodology and Results for Typical Aerostat Shapes In order for the solar powered aerostat to produce the minimum requirement of 100kW average rating, an initial sizing procedure was done to see how much area is needed and the weight of the panels. Again, the area required is a function of the PV cell efficiency, solar irradiation, and the weight is dependent on the density of the solar panels. These factors range greatly from one type of solar cell technology to the other. Table 24 shows the results of the calculations using three different materials lightweight amorphous silicon (A-Si) triple junction cells, triple junction with monolithic diode high efficiency cells, , and a more conventional mono-crystalline solar panels. The first two are relatively new technologies compared to the mono-crystalline cells, which are readily available and used in many household installations. The A-Si option requires much more area, but due to its low density it is significantly lighter than the latter option. However, the monolithic diode has a much higher efficiency so it requires less area to produce the same amount of energy, but it is high efficiency comes with a weight

71

penalty. From here, the sizing of the aerostat body can be calculated using initial sizing procedures discussed later in this section. (54) The requirement is to create a solar power system that yields at least 100kW power output comparable to the Altaeros design in order to provide a fair analysis. Initial sizing was carried out for two common aerostat shapes with each different type of solar panel, a GNVR shape and an NPL shape, at 100 meters altitude. After calculating the necessary area and associated weight, the preliminary sizing methodology and Excel formulation of Gupta was used with a payload weight set equivalent to the solar cell weight. It is clear that for the very light weight amorphous silicon cells, the aerostat is much too small to generate a large amount of power, as the area associated with the design to lift the given payload is a full order of magnitude lower than the necessary 100kW value. Utilizing many smaller aerostats in a larger array would probably be the best way to generate the necessary power using these cells. When compared to the amorphous silicon cells, the monolithic diode cells require much less area to obtain the 100kW power rating, but the associated weight of such a system is much larger. The preliminary sizing analysis shows that the volume required to lift the heavier payload is significantly larger than in the amorphous silicon case, but the surface area is still insufficient. The mono-crystalline option requires an area midway between the previous two options to generate the 100kW, but due to their much higher weight, the volume required is impractically large. The area condition is more than satisfied, although the extremely large volume proves prohibitive. (18) (54) In order to attain the desired power output, the area requirement must be satisfied without overly exceeding the volume required for the given weight. To achieve this, a shape with a larger surface area to volume ratio should be explored. One such shape is an oblate spheroid; it provides a larger surface area for a similar internal volume. The following section provides the first results obtained when optimizing for a minimum volume oblate spheroid aerostat. Table 24 includes analysis of three different types of solar panels based on minimum area requirement the ratio of the total area of the aerostat to the area occupied by PV cells is approximately 2.00 . The code developed based on the methodology above, outputs values for two different gases and envelope materials used, to portray effects of permeability. Such outputs are: the minimum volume required for minimum area, the time afloat as a function of these conditions, the excess lift needed, and characteristics of the shape of the aerostat. Thus, the calculations described in the above method yielded the values in Table 24. The figures representing the oblate spheroid shape in Table 24 were generated using the MATLAB function ellipsoid() with the earlier calculated semi-major, and semi-minor axes of the elliptical cross section from initial sizing, and with shape center at (0,0,0) as inputs. It should be noted that for an ellipsoid to become an oblate spheroid, the radius of the circular section is equal to the semimajor axis of the elliptical profile.

72

Table 24: Oblate spheroid analysis for different types of PV cells

The shows that the thickness to chord ratio for the shape-optimized feasible 100 kW concept is 5.5% and the plausible type of PV cells one can use based on weight-to-efficiency ratio is the triple junction monolithic diode type (28% efficiency). Even though the flatness of a spheroid is desirable for improving PV cell efficiency, taking manufacturability into account, such a model would be difficult to successfully deploy. As a result 10%, 15%, and 20% thickness to chord ratios were considered to account for such issues and serve as a more effective means to comparing the sizing and scaling-up of the different PV-cell aerostats. However, these configurations seemed hard to manufacture as well, while it was noticed that the volume of helium caused by the t:c modifications yielded many years of endurance. As a result, the methodology was revised and modified slightly to reflect ease of manufacturability, and realistic endurance of the envelope portion of the system while meeting power requirements needed by the up-scaling process. Revised Sizing Revised Envelope Sizing Methodology For the conceptual design configuration, the following method was implemented in order to analyze its properties. It should be noted that after determining the dimensions of the aerostat based on the area required to yield a particular power output the lift capacity required to keep the aerostat afloat for 36 months was calculated. This value is taken as a fraction (%) of the 90% He lifting capacity (approximately 1.05 kg/m3) and rounded to the closest integer (%) which produces the extra time afloat (Table 24) a number greater than the minimum requirement of 36 months. The following revised methodology was then carried for a final time, with the adjusted lift capacities required (see Fraction of He Lift Capacity Required column in Table 25, Table 26, and Table 27 in Revised Envelope Sizing Results) for 25%, 33%, and 50% thickness to chord ratios and power outputs ranging from 100 kW to 1MW
73

For analysis purposes, the semi-minor axis to semi-major axis ratio is referenced later in the report as the thickness to chord ratio. The volume and surface area used in determining the aerostat dimensions for a particular power output are described by the following equations: ( ) ( (( ) ( )))

The initial sizing and profile characteristics calculations were carried as follows based on the geometrical characteristics found above. The length of the tether can be related by:

where H is the design altitude of the aerostat. As one can later see in the Tether and Mooring Section, three elastic Dynema SK75 are used for load bearing and stability. The extra 30% of length is implemented to account for blowby wind. The fourth tether is nonloadbearing and contains the power conducting elements of the system. Thus, the power conducting cable is 50% longer than the required altitude H. Accordingly, the overall weight of the tether is a function of the unit weight of each tether and the design length. Thus, the overall weight of the tether ( ) is found by: Then if the area density of the PV cells is introduced as cells is introduced as , then the weight of the aerostat ( ( and the surface coverage fraction of the PV )can be related by: )

in which is the density of the envelope material. Note that 1.33 is applied as factor to account for overlaps and patches in the envelope material as well as internal webbing with. The internal webbing material tensile strength must withstand internal pressure stresses. At this stage, in order for the aerostat to maintain its shape (using webbing) the internal pressure needed is estimated according to the following expression: . It should be noted that if there is an internal rigid structure present instead of webbing, the pressure need not be as high. Subsequently the overall weight of the design ( ) can be calculated ( )

Defining the lift capacity of a gas as (Helium lift capacity taken as 1.058 kg/m3 for 100% He envelope volume), the overall lift capability of the design becomes,

where
[

for which Tmin is 36 months and the Helium percent of the enclosed volume is the difference between the total volume of the aerostat and the volume occupied by air air weight being included in payload estimations. The He : air ratio can be found through variations by 1%, until the first lcmin value that is less
74

than 1.058 kg/m3. From the lift capability it is then possible to determine endurance properties of the design. Given a material permeability, the rate of gas leaking (dV) from the envelope is found through:

where perm is the permeability of the material. After the optimum Helium to air percentage of aerostat volume is determined and reasonably rounded to closest percent value. The excess lift provided by the aerostat is then:
( )

Thus the operation endurance (in days) that the aerostat can maintain an altitude down to the minimum operational altitude is ( )

lc is the lift capacity of the Helium lifting gas multiplied by the fraction needed for a particular volume to lift the payload, and dV is the leakage rate of the gas a function of aerostat area and permeability of the envelope material. The variable T stands for the amount of time that the aerostat can maintain at least the minimum operational altitude. Additionally, the excess lift is taken at a minimum of 5% of the required lift force in order to maintain tension in the tether for any configuration. Although for weight estimation purposes, the internal webbing of the aerostat along with overlapping patches of the envelope were assumed to account for 30% of the envelope weight, the crosssectional area of the oblate spheroid is determined to later serve as input for determining the internal structure specifications:

where a is the semi-major axis and c is the semi-minor axis of the elliptical cross section, found from the required surface area, and endurance-based volume. A more realistic thickness to chord ratio (25%, 33%, and 50%) is implemented, which raises the curvature of the spheroid when compared to the previous report specifications, and thus the volume of the aerostat. In absence of information, in an attempt to reduce the PV cell efficiency loss due to the effects of the curvature of the aerostat, it was decided that only 75% of the top surface of the aerostat (center) is to be considered for mounting solar panels, which was applied to the 25% t:c balloon ( is then 37.5%). Consequently, since increasing the thickness to chord ratio accentuates the curvature of the aerostat, 70% and 65% of the upper half area of the aerostat were studied for PV cell mounting for the 33% and 50% t:c respectively. Additionally, with the volume found, the resulting extra static lift force added to the dynamic lift caused by the maximum operational wind conditions, dictates the sizing of the mooring system previously described With the sizing finalized, the drag analysis on the oblate spheroid according to Hoerner for laminar flows can be estimated as a function of Cf (assumed to be 0.005 for Vectran) (40): ( ) ( ) ( )

The zero lift drag coefficient is used in drag calculations assuming that ideally the three tethers and the associated winching systems will keep the aerostat at a 0 angle of attack. For supercritical Re

75

numbers Hoerners CDo graph for different thickness to chord ratios for ellipsoidal bodies is shown in Figure 57

Figure 57: Drag Coefficient of Ellipsoidal Bodies at Subcritical and Transition Re (30) and can be determined according to the formula: [ ( ) ( ) ]

using a Cf of 0.004 (40). Computing drag as a force acting on the envelope can be done as follows knowing CDo for the aerostat:

where is 1.29 kg/m3, U is 8 m/s at the chose altitude, and V2/3 volumetric reference area. Consequently, the drag and lift analysis can then be used for finding the tension in the tethers. The conclusions clearly stated by the drag equations, are that the least t:c ratio that can be manufactured is the most efficient. It was decided that the lowest thickness-to-chord-ratio driven by manufacturability is 25%; however, 33%, and 50% t:c were analyzed as well for its buoyancy and gas cost saving purposes. Revised Envelope Sizing Results The tables that follow present the characteristics required by the aerostat power system to yield 100 kW, 250kW, 500 kW, and 1 MW based on the methodology outlined in the previous section. The lifting capacity fraction is rounded to the closest percent, with a minimum of no less than 5% (see Table 25, Table 26, and Table 27 for the 500kW and 1 MW configurations) which gives the extra time afloat (greater than 3 years). Note that Fraction of He Lift Capacity Required refers to the lift capacity required to lift the payload without taking into account air weight, while % Volume of He Required is the lift required to lift the payload and the air volume percentage in the envelope.

76

Table 25: Solar balloon specifications for 25% t:c, with 37.5% of area covered by PV cells
Fraction of He Lift Capacity Required 20% 13% 9% 6%

25% t:c 100kW 250kW 500kW 1MW

Area (m2) 3970 9922 19843 39683

Volume (m3) 14476 55960 152962 447677

CS Area (m2) 452 1119 2165 4477

Total Aerostat Mass (kg) 8999 30713 86262 251020

% Excess Lift 10% 9% 13% 10%

Excess Lift Force (N) 9385 30812 122203 278898

% Volume of He Required 65% 62% 61% 59%

Time Afloat (months) 37 49 98 111

Table 26: Solar balloon specifications for 33% t:c, with 35% of area covered by PV cells
Excess Lift Force (N) 9800 32385 58430 164926 Fraction of He Lift Capacity 16% 10% 7% 5% % Volume of He Required 63% 60% 58% 57%

33% t:c 100kW 250kW 500kW 1MW

Area (m2) 4129 10318 20636 41270

Volume (m3) 18506 71681 198214 573445

CS Area (m2) 591 1453 2859 5812

Total Aerostat Mass (kg) 10817 40382 112200 322090

% Excess Lift 8% 8% 5% 5%

Time Afloat (months) 38 50 45 63

Table 27: Solar balloon specifications for 50% t:c, with 32.5% of area covered by PV cells
Excess Lift Force (N) 14560 42542 68070 122968 Fraction of He Lift Capacity 13% 8% 5% 5% % Volume of He Required 62% 59% 57% 56%

50% t:c 100kW 250kW 500kW 1MW

Area (m2) 4288 10715 21430 42857

Volume (m3) 22301 92363 261799 718378

CS Area (m2) 760 1979 3927 7697

Total Aerostat Mass (kg) 13144 51717 150940 413090

% Excess Lift 10% 8% 4% 3%

Time Afloat (months) 54 63 50 45

The general shape of the aerostat as taken at a 1:2 thickness to chord ratio for 100kW power output is shown in Figure 58.

77

20 10 0 -10 -20 20 15

10

-5

-10

-15

-20

-20

-15

-10

-5

10

15

20

Figure 58: Triple Junction with Monolithic Diode Balloon: 50% t:c, 100kW, 32.5% of area covered by PV cells (axes in m) Thus, the sizing and scaling of the 100 kW model to accommodate up to 1 MW of power (data points shown for 100 kW, 250 kW, 500 kW, 1 MW, read from left to right) can be illustrated as shown in Figure 59, where the reference area and reference volume are the values for the 25% t:c, 100 kW model: 3970 m2 and 14476 m3.

Figure 59: Sizing and Scaling Solar Balloon Concept from 100 kW to 1 MW (data points shown for
100kW, 250kW, 500kW, 1 MW)

78

Mooring and Tether After the general tether analysis conducted for all aerostat concepts, it was determined that for higher stability required by a spherical inflatable shape multiple tethers should be investigated. This is the case for the solar balloon aerostat. Multi-tether analysis Studies carried out in the past years on tethered spherical aerostats result in a typical number of three tethers for mooring purposes. Results demonstrate that a tethered sphere will oscillate strongly at a saturation amplitude of close to two diameters peak-to-peak in a steady fluid flow. The oscillations induce an increase in drag and tether angle on the order of around 100% over what is predicted using steady drag measurements. The in-line oscillations vibrate at twice the frequency of the transverse motion (30) Typically, it has been observed that a three tether system is often used for stability of a spherical aerostat, to counteract the tendency of the balloon to move in a figure eight motion, and also to counteract the increase in drag resulting from that motion (55) From data gathered in the general Tether section of this report it was concluded that the three loadbearing tethers are to be made of Dynema SK75 at 130 m long, while the power-conducting cable used by Magenn would be nonloadbearing (fourth) tether at 150 m in length. Assuming an R:Rw of 0.4 , with a minimum angle from the ground of 22.3, the maximum achievable altitude is R. While cases like the 100 kW and 250 kW models can be deployed at 100 m, massive models (500 kW, and 1 MW models) have to be deployed at a maximum allowable altitude assumed to be 700 m according to Altaeros (55) (16). Thus the length required by each tension tether can be found through simple trigonometry. However, the wind speeds are higher at 700 m altitude, and stability may become a relatively bigger issue, although the multi-tethered system will counteract most of these problems. The drag analysis on the tether was carried out in the same fashion as the single tethered aerostats and summed for each cables. The sizing of the envelope methodology accounts for the multiple tether weights, while the excess lift produced needs to be counteracted by the four winches with swivel type mooring masts in Figure 60

79

Winch 1 Power conducting cable Winch 4

Winch 3

Winch 2

Figure 60: Multi-tethered Solar Power Aerostat (adaptation) (55)

Solar Aerostat Cost General assumptions Although the mobility of the designs was less of a factor, the balloons are not installed such that they cannot move. Unlike traditional PV, the balloons can be transported to other sites in relatively short amounts of time. The systems are also designed to be off the grid, reducing installation costs significantly. Therefore, estimating prices for installation proved negligible, and the prices were included in the operating and maintenance cost. For standard solar panels, operating and maintenance costs can be assumed at a rate of $0.01 to $0.03 per watt with a higher installation cost. The balloons will be expected to handle the elements, as theyre constructed of fabric, and will therefore require more maintenance. For this reason, a yearly cost of 1% of the capital cost is assumed for operating and maintenance. (3) (4) For equation 1, a discount rate of 5% was used, with a lifetime of 20 years. Typical PV cells last for at least 20 years, and they are the bulk of the cost for this design. (5) Replacing and fixing the balloon is factored into the O/M costs. Discount rates range between 5% and 10%. A 10% discount rate would make the project too expensive, and was therefore not evaluated.

Cost/Pricing Analysis Table 28 contains the reference costs used for each solar balloon design, including the labor and manufacturing cost:
80

Table 28: Reference Cost of Solar Aerostat System Components


Component Helium ($/m^3) Envelope ($/m^2) Power ($/W) Tether ($/m) Webbing ($/m^2) Confluence line ($/m) Cost $0.28 $25.00 $3.00 $12.00 $7.50 $3.33

Using the general manufacturing assumptions taken for all models with the specific assumptions for the balloon already described, the capital prices for the solar balloons, including lifetime helium usage, are specified in Table 29. Table 29: Total Investment Cost (in thousands of dollars)

Power Rating 100 kW

Power 300

Winch 80

t/c 25% 33% 50% 25%

Helium 19.7 23.9 21.1 57.3 69.8 73.4 154.1 203.8 242.3 233.9 440.2 713.4

Envelope 99.3 103.2 107.2 248.0 258.0 267.9 496.0 515.9 535.8 992.1 1031.8 1071.4

Webbing 29.8 31.0 32.2 74.4 77.4 80.3 149.0 155.0 161.0 298.0 309.5 321.4

TOTAL 535.3 544.6 547.0 1246.3 1271.7 1288.1 2491.5 2567.0 2631.3 4876.0 5133.9 5458.7

250 kW

750

110

33% 50% 25%

500 kW

1,500

160

33% 50% 25%

1000 kW

3,000

320

33% 50%

From this analysis it is evident that as the power rating increases, the helium increases due to the increased volume. Aside from the 100 kW rating balloon, this trend also holds for the thickness to chord ratios. In that as the thickness to chord ratio increases, the amount of needed gas over the systems lifetime increases. A design consideration taken into account was the reduction in the amount of helium gas, cutting the ratio of helium to air from 100:0 to more considerable ratios. This was done to reduce cost, as filling the balloons with 100% helium produced more excess lift than needed. The amount of helium needed over the lifecycle of the balloon with the time afloat and percent Helium volume are provided in Table 30.

81

Table 30: Helium Effects on Solar Aerostats Lifecycle


Lifecycle Volume of Helium (m3) 70,444.70 85,294.80 75,279.60 204,634.50 249,454.30 262,095.40 550,329.60 728,119.20 865,522.80 835,232.30 1,572,085.10 2,547,893.10 Time afloat (days) 1126.2 1156.6 1643.6 1491.4 1521.8 1917.5 1491.4 1369.7 1521.8 3378.5 1917.5 1369.7 Percent Helium 65% 63% 62% 62% 60% 59% 61% 58% 57% 59% 57% 56%

Power Rating (kW) 100

t/c 25% 33% 50% 25%

250

33% 50% 25%

500

33% 50% 25%

1000

33% 50%

The higher power-rating balloons required substantially more envelope area to carry all of the solar panels. This rate at which the area grew was faster than the volume of helium required, as each balloon had different helium:air ratios, and the larger balloons had lower ratios. The cost of the webbing was assumed at 30% of the envelope cost. The winching and mooring system is designed such that the balloons can be moved to a different area off the grid relatively quickly. However, with more tethers being used in this design, more winches are required. It can be safely assumed that the four winches will still add up to the original, giving the same original winching price. Each individual winch in this designed is assumed to be one quarter the price of the standard winches. The most expensive and important aspect of the balloons are the PV cells. Standard pricing for the particular cells were taken at $3.00 per watt, which includes labor and manufacturing. This pricing scheme was developed by looking at a wide range of solar plants and determining their cost per wattage; the exact PV cells which were used and their costs per wattage on a small scale were then determined, with an added assumption of 30% reduction in material cost through mass-purchase. In the next section the projected costs were put through Equation (5.5) as an algorithm, and the projected cost of electricity over the lifecycle of the balloon was calculated. These costs can then be compared with traditional methods of generating power, as well as current PV schemes already in practice to determine if such a product is cost-feasible. Cost of Electricity Generation As stated previously, the levelised electricity generation cost can be determined with Equation (5.5). Each t/c value was analyzed at four output ratings, 100, 250, 500 and 1000 kW. The lifetime electricity generation was calculated assuming a lifetime of 20 years, with 4 hours of collection per day, or a capacity factor of 16.67%. The lifetime electricity generation is calculated using : ( ) ( ) ( )
82

where: Lifetime electricity is the total estimated electricity generated by the balloon measured in kW-h Power rating is the maximum power output of the device measured in kW The capacity factor is the ratio of actual power generated divided by its full operating potential Lifetime is the lifecycle of the balloon measured in days

The resultant life, yearly and daily electricity generation for each output rating is listed in Table 31 assuming a capacity factor of 0.1667 and a lifecycle of 20 years.

Table 31: Life, Yearly and Daily Electricity


Power Rating (kW) 100 250 500 1000 Daily Electricity (kW-h) 400 1,000 2,000 4,000 Yearly Electricity (MW-h) 146.1 365.3 730.5 1,461.0 Lifecycle Electricity (MW-h) 2,922 7,305 14,610 29,220

Taking the yearly electricity generation, the total investment costs and operation and maintenance costs and using the algorithm in Equation (5.5), the total lifetime costs are listed in Table 32.

Table 32: Solar Concepts Total Lifetime Cost


Power Rating (kW) 100 t/c 25% 33% 50% 25% 33% 50% 25% 33% 50% 25% 33% 50% LEC ($/MW-h) 336.2 342.0 343.5 313.1 319.5 323.6 309.7 319.2 327.3 304.6 320.8 341.2 O/M ($/MW-h) 36.6 LEC TOTAL ($/MW-h) 372.8 378.7 380.2 347.2 353.6 357.7 343.4 352.9 361.0 337.8 354.0 374.4

250

34.1

500

33.8

1000

33.2

Conclusions Conceptually speaking, one would expect the price of the design to decrease as the power rating increases. This is precisely the trend that is observed. The 100 kW design is by far the most expensive while the 1000 kW design is the least expensive, in terms of lifecycle cost.
83

Different t/c values were taken into account to determine if the difference in helium made up for the difference in envelope area. The amount of helium relative to the air decreases with an increasing t/c, thus it is possible that less helium would have been used, and that this lower amount of helium would compensate for the increase in area. However, this was determinedly not the case, and the lowest possible t/c value is also the most advantageous. This could change in the future as the price of helium is expected to increase dramatically, and different excess lift calculations can be made to reduce the needed lifting gas overall. As can be seen in example Table 29, the PV cells make up the bulk of the cost. It is therefore the most influential aspect of the design. If the cost of the PV cells are cut in half, a 25% reduction in levelized energy cost is reached, or approximately ten cents per kW-h. This was observed regardless of the power output of the balloon. The lifting gas was also a key driver, but not as important. If the cost of helium was doubled, it would result in a 5% increase in levelized energy cost, or approximately $0.004 per kW-h. While this increase is not negligible, it does not influence the design to the degree that the PV cells do. The envelope cost was very influential in the final cost, not only because it had the second highest percentage of the overall design cost, but because the internal webbing was linearly dependent on the envelope itself. Therefore an increase in envelope cost results in a higher webbing cost as well. Doubling the cost of the envelope increased the LEC by 23%, or 9 cents per kW-h. The winching systems importance was wholly dependent on which design was chosen. It was a very significant aspect of the 100 kW system, but was much less important on the 1 MW balloon as it took up a smaller percentage of overall costs. The tether and confluence lines were fairly negligible overall, regardless of the power output. Even with a fivefold increase, the LEC increased by 0.5%. With these results, the cost drivers were able to be analyzed and listed in Table 33. Table 33: Solar Aerostat Cost Drivers
Rank 1 2 3 4 5 Component PV Cells and energy generation Envelope Lifting gas Winching system Confluence lines and tether

Though the winching system influences the cost of the design to a greater degree than the gas, it is reasonable to assume that the cost of helium will rise at a faster rate than the manufacturing and material cost for the winch. This is why the gas was seen as a more important design driver than the winch. Even if its impractical in size dimensions, the 1 MW system was by far the least expensive, and is the most viable in terms of cost alone. For comparisons sake, its viability on the cost front was compared below in the General Conclusions section.

Solar Concept Conclusions According to the assumptions made in previous sections of the report and the cost conclusions, it was determined that the most efficient solar system is the 1 MW, 25% t:c, 150 diameter model. The solar power aerostat manufacturability can be inferred from existing airship dimensions. In that aspect
84

Hindeburg airship had a chord of 245 m (56). Despite the fact that at first glance, the result values render the most efficient model impractical, it is an attractive concept in places where the land is scarce and building a terrestrial solar power station is not an option. This aerostat system can provide power while also becoming a large protective roof against solar radiation (57).

FINAL CONCLUSIONS Figure 23 contains the minimum, average and maximum cost for different forms of energy by todays standards. Table 34 contains the final designated system, on a cost-based analysis alone, for each design system. Table 34: Finalized designs for comparison (in thousands of dollars)
Design DAWT Solar Balloon Magenn Power Rating (kW) 100 1,000 1,000 Capacity Factor 0.450 0.167 0.450 Lifecycle (years) 20 20 15 LEC ($/MW-h) 107.9 304.6 82.2

When compared with the maximum amount on coal, Magenn and the DAWT can compete with the energy prices of today, if the assumptions made were correct and within a reasonable margin of error. Though they can compete, coal is on an unfair playing field both because its subsidized, and because its negative externalities arent paid for. It is therefore necessary to tax the carbon produced by dirty sources like coal and natural gas. Its been estimated that $12 per tonne of carbon dioxide is required (58). If the DAWT and Magenn operate as calculated, they can compete with on-shore wind power. The Magenn concept is the most cost effective solution, by far, whereas the solar balloon concept seems too expensive to be feasible. The solar balloons can compete with the maximum side of the curve for solar PV, but this is not a fair comparison as this was the minimum cost for all of the solar balloons. It is from these results, on a cost basis only, that Magenn 1 MW envelope is the most cost effective solution.

85

References 1. Independent Statistics and Analysis. US Energy Information Administration. [Online] http://www.eia.doe.gov/. 2. Group, Energy Watch. Coal: Resources and Future Production. [Online] March 2007. http://www.energywatchgroup.org/fileadmin/global/pdf/EWG_Report_Coal_10-07-2007ms.pdf. 3. Rogers, Alison. Wind Power: Are Vertical Axis Turbines Better? Mother Earth News. [Online] Mar 2008. [Cited: Oct 12, 2010.] http://www.motherearthnews.com/Renewable-Energy/2008-02-01/WindPower-Horizontal-and-Vertical-Axis-Wind-Turbines.aspx?page=5. 4. Wind Research: Wind Systems Integration. National Renewable Energy Laboratory. [Online] U.S. Department of Energy, August 24, 2010. http://www.nrel.gov/wind/systemsintegration/. 5. Wind Powering America: 80-Meter Wind Maps and Wind Resource Potential. Wind and Water Power Program: Wind Powering America. [Online] [Cited: 2010 11-Nov.] http://www.windpoweringamerica.gov/wind_maps.asp. 6. Pant, R.S. Introduction to Aerostats. Bombay : Indian Institute of Technology. 7. Allsop Helikites Limited. [Online] Allsop Helikites Limited, 2008. http://www.helikites.com/bird/index.html. 8. Foreman, K, Gilbert, B and Oman, R. Diffuser Augmentation of Wind Turbines. Solar Energy. 1978, Vol. 20, pp. 305-311. 9. Phillips, D. G., P. J. Richards, and R. G. Flay. DIFFUSER DEVELOPMENT FOR A DIFFUSER AUGMENTED WIND TURBINE USING COMPUTATIONAL FLUID DYNAMICS. The University of Auckland New Zealand. [Online] 10. Widnall, Sheila. Potential Flow Calculations of Axisymmetric Ducted Wind Turbines. Boston : Massachusetts Institute of Technology, 2009. 11. MARS. Magenn: Wind Power Anywhere. [Online] Magenn Power Inc. [Cited: November 19, 2010.] http://www.magenn.com. 12. Magenn Wind Power Technologies. Rivard, Pierre. Blacksburg : s.n., 16 Nov 2010. 13. Solar Energy Fact Sheets. Solar Developments. [Online] [Cited: April 14, 2011.] http://www.solardev.com/SEIA-lightworld.php. 14. Chino, Mike. Solar Balloons: Sun Hope Renewable Energy. Inhabitat. [Online] [Cited: November 19, 2010.] http://www.inhabitat.com/2008/04/10/sunhope-solar-balloons. 15. Cool Earth Solar. Cool Earth. [Online] [Cited: April 14, 2011.] <http://www.coolearthsolar.com/>. 16. Glass, B. Concept Note for Airborne Wind Turbine. Private Communication : Altaeros Energies, July 2010. 17. G. S. Aglietti, T. Markvart, A. R. Tatnall and S. J. Walker. Solar Power Generation Using High Altitude Platforms: Feasibility and Viability. 18. Pant, R. Conceptual Design Phase Study Report of Program on Airship Design and Development. Indian Institute of Technology Bombay : Department of Aerospace Engineering, 2002. 19. A Novel Concept for Stratospheric Communications and Surveillance: Star Light.". Chu, Adam, et al., et al. s.l. : American Institute of Aeronautics and Astronautics, 2006. 20. Khoury G. A., Gillet J. D., Eds. Airship Technology. s.l. : Cambridge University Press. ISBN 0 521 430 737. 21. Pattinson, John. Manufacture, Flight and Operation HALE airship. Shropshire, UK : Lindstrand Technologies Ltd. 22. Harrison, Karl. Chemistry, Structures & 3D Molecules. N.p. [Online] March 2007. [Cited: Nov 14, 2010.] http://www.3dchem.com/molecules.asp?ID=326. 23. Tedlar PVF Fils. Dupont. [Online] Dupont. [Cited: April 14, 2011.] <http://www2.dupont.com/Tedlar_PVF_Film/en_US/assets/downloads/pdf/Tedlar_GeneralProperties.pdf >.. 24. Miller, Jonathan I. and Meyer, Nahon. Analysis and Design of Robust Helium Aerostats. Journal of Aircraft. 2007.
86

25. TCOM: Builder of World Class LTA Systems. TCOM. [Online] 2009. [Cited: April 14, 2011.] http://www.tcomlp.com/. 26. Raina, A.A. and Bhandari, K. M. Conceptual Design Of A High Altitude Aerostat For Study Of Snow Pattern. ITT Bombay. Manali, India : s.n., 2009. Proceedings of International Symposium on Snow & Avalanches. 27. Sullivan, Daniel. Zeppelin Eureka Heads To Los Angeles & Orange County. Daggle Danny Sullivan's Personal Blog. [Online] [Cited: Nov 2010, 19.] <http://daggle.com/zeppelin-eureka-heads-laand-oc-672. 28. The future of LTA systems. Lockheed Martin. [Online] [Cited: Nov 19, 2010.] http://www.lockheedmartin.com/data/assets/ms2/pdf/AerostatSystems.pdf. 29. Wichard Stainless Steel Mooring Swivel. Green Boat Stuff. [Online] [Cited: April 2011, 18.] http://www.greenboatstuff.com/wiststmosw.html. 30. Zhao, Xiaohua. Statics and Dynamics Simulation of a Multi- Tethered Aerostat System . Nanjing : A Thesis Submitted in Partial Fulfillment of the Requirements for the Degree of MASTER OF APPLIED SCIENCE in the Department of Mechanical Engineering, Nanjing University of Aeronautics & Astronautics. 31. Dyneema. Midas Coporation. [Online] [Cited: Nov 19, 2010.] http://www.midasglove.com/english/viewforum.php?f=47. 32. Three-Dimensionally Braided Carbon FiberEpoxy Composites, A New Type of Material for Osteosynthesis Devices. I. Mechanical Properties and Moisture Absorption Behavior. Wan, Y. Z., Y. L. Wang, G. X. Chen, and K. Y. Han. 2001, Journal of Applied Polymer Science. 33. Carbon Fiber Braided Sleeving. Cable Ties and Sleeving - Cable Management System and Wiring Accessories. [Online] [Cited: Nov 19, 2010.] http://www.cabletiesandmore.com/carbon-fibersleeving.php. 34. Forrester, Christopher J. von Alt and Ned C. The Design of Small Diameter Neutrally Buoyant Electro- Optic Tether Cables for Maximum Power Transfer. [Online] [Cited: Nov 19, 2010.] <http://ieeexplore.ieee.org/stamp/stamp.jsp?tp=&arnumber=364089&isnumber=8343. 35. Beach. Lightning Hardened Tether Cable and an Aerostat Tethered to a Mooring System. 4842221 United States. 36. Grounding Instructions for a TV Antenna. EHow.com. [Online] http://www.ehow.com/how_6720916_grounding-instructions-tv-antenna.html. 37. Rivard, P. Private Communiation. April 14, 2011. 38. Wright, J.B. Computer Programs for Tethered Balloon System Design and Performance Evaluation. Hanscom Air Force Base : Air Force Geophysics Laboratory, August 1976. AFGL-TR-77-0203. 39. Harper, C.A. Modern Plastics Handbook. New York : McGraw-Hill, 2000. 40. Hoerner, Sighard. Fluid Dynamics Drag. Bakersfield, CA : Hoerner Fluid Dynamics, 1993. 41. Howard, Alistair. Experimental Characterization and Simulation of a Tethered Aerostat with Controllable Tail Fins. s.l. : Department of Mechanical Engineering McGill University, 1987. 42. Morgan, Jason. Comparing Energy Costs of Nuclear, Coal, Gas, Wind, and Solar. Nuclear Fissionary. [Online] NuclearFissionary.com, April 2, 2010. http://nuclearfissionary.com/2010/04/02/comparing-energy-costs-of-nuclear-coal-gas-wind-and-solar/. 43. Broehl, Jesse. A New Chapter Begins for Concentrated Solar Powe. Renewable Energy World. [Online] February 11, 2006. [Cited: April 2011, 18.] http://www.renewableenergyworld.com/rea/news/article/2006/02/a-new-chapter-begins-for-concentratedsolar-power-43336. 44. Fletcher, Herman S. Experimental Investigation of Lift, Drag, and Pitching Moment of Five Annular Airfoils. Langley Field, VA : NACA, 1957. 45. Getting clean energy into high gear at MIT (photos). ZDNet. [Online] CBS Interactive, 2010. http://www.zdnet.com/photos/getting-clean-energy-into-high-gear-at-mit-photos/6199965. 46. Surface of Revolution. Wolfram Mathworld. [Online] http://mathworld.wolfram.com/SurfaceofRevolution.html.
87

47. Volume of Revolution. Wolfram Mathworld. [Online] http://mathworld.wolfram.com/VolumeofRevolution.html. 48. Devenport, William J. Vortex Panel Method. [Online] http://www.engapplets.vt.edu/fluids/vpm/vpmex.html. 49. Ernst A. van Nierop, Silas Alben, and Michael P. Brenner. How Bumps on Whale Flippers Delay Stall: An Aerodynamic Model. s.l. : Physical Review Letters, 2008. 50. Foreman, K. M. Preliminary Design and Economic Investigations of Diffuser Augmented Wind Turbines (DAWT). New York : Grumman Aerospace Corporation, 1981. 51. S. Gupta, S. Malik, M. Kumar, and A.R. Laskar. Envelope Details for Demo Airship. s.l. : ADRDE, PADD Project, 2002. 52. High-Lift Aerodyanmics. Smith, A.M.O. 6, s.l. : Journal Of Aircraft, 1975, Vol. 12. 53. Bertin, John J and Russel, Cummings M. Aerodynamics For Engineers 5th Edition. Upper Saddle River : Pearson Prentice Hall, 2009. 54. Aglietti, G. S., S. Redi, and A. R. Tatnall. Aerostat for Solar Power Generation. Sciyo. [Online] [Cited: Nov 14, 2010.] <http//:www.sciyo.com>. 55. Casey Lambert, Aaron Saunders, Curran Crawford and Meyer Nahon. Design of a One-Third Scale Multi-Tethered Aerostat System for Precise Positioning of a Radio Telescope Receiver. Dept. of Mechanical Engineering , McGill University. Montreal Quebec, Canada, H3A 2K6 : s.n. 56. Grossman, Daniel. Hindenburg Statistics. Airships: The Hindenburg and other Zeppelins. [Online] 2010. [Cited: April 18, 2011.] http://www.airships.net/hindenburg/size-speed. 57. Pant., Rajkumar. Private e-mail communication. April 2011. 58. Tideman, Nicolaus. Pricing Externalities. Blacksburg : s.n., 2009. Economic Report. 59. Frequently Asked Questions. Next Era Energy Resources. [Online] Next Era Energy Resources, LLC, 2011. [Cited: February 1, 2011.] http://www.nexteraenergyresources.com/content/where/portfolio/wind/faq.shtml. 60. Cape Wind: America's First Offshore Wind Farm on Nantucket Sound. Cape Wind Associates, LLC. [Online] Cape Wind Associates, LLC, 2010. [Cited: January 16, 2011.] http://www.capewind.org/. 61. Portman, Don. The Nocturnal Low-Level Jet. Windwisdom. [Online] University of Michigan, 2004. [Cited: January 10, 2011.] http://www.windwisdom.net/nllj.htm. 62. FloDesign Wind Turbine. PESWiki. [Online] Pure Energy Systems. [Cited: November 2010, 2010.] http://peswiki.com/index.php/Directory:FloDesign_Wind_Turbine. 63. Karasudani, Takashi, and Ohya, Yuji. A Shrouded Wind Turbine Generating High Output Power with Wind-lens Technology. Energies. March 31, 2010, Vol. 3, pp. 634 - 649. 64. Wind Turbine Power Calculations. s.l. : The Royal Academy of Engineering. 65. Gipe, Paul. Ducted or Augmented Turbines. Wind-Works.org. [Online] October 12, 2009. [Cited: February 16, 2011.] http://www.wind-works.org/SmallTurbines/DuctedorAugmentedTurbines.html. 66. Pattinson, John. Manufacture, Flight and Operation HALE airship. Shropshire : Lindstrand Technologies Ltd. 67. Stern, S. Alexander, Mohr, Paul H., and Gareis, Paul J. Recovery of Helium. 3,426,449 United States, April 19, 1966. 68. Weller, Sol W. Recovery of Light Elemental Gases. 2,540,152 United States, Feb 6, 1951. 69. Design, Fabrication and Operation of Remotely Controlled Airships in India. Gawale, Amol, Amool Raina, Rajkumar Pant, and Yogendra Jahagirdar. 2008, INTERNATIONAL CONGRESS OF THE AERONAUTICAL SCIENCES. 70. Metal, Plastic, and Ceramic Search Index. Online Materials Information Resource - MatWeb. [Online] [Cited: Nov 19, 2010.] http://www.matweb.com/. 71. OPGW Product. MadeInChina.com. [Online] http://www.made-inchina.com/showroom/sisi2718/product-detailafPEnrpWxQJo/China-OPGW.html. 72. Mason, W.H. Drag: An Introduction. [book auth.] Grumman, Rino Roman R. Hendrickson. Configuration Aerodynamics. 2006.

88

73. Nagabhushan, Luis Valera and Bellur. DESIGN TRENDS AND GLOBAL DEVELOPMENTS IN MODERN LTA VEHICLES. Saint Louis, Missouri : Saint Louis University, 2002.

89

Potrebbero piacerti anche