Sei sulla pagina 1di 9

Proceedings of the ASME 2010 International Mechanical Engineering Congress & Exposition IMECE2010 November 12-18, 2010, Vancouver,

British Columbia, Canada

IMECE2010-38395
A COMPARATIVE ANALYSIS OF WIND PROPULSION SYSTEMS FOR OCEANGOING VESSELS

Benjamin H. Gully, MSME Graduate Research Assistant Mechanical Engineering Department The University of Texas at Austin

Dr. Michael E. Webber Assistant Professor Mechanical Engineering Department The University of Texas at Austin

Dr. Carolyn C. Seepersad Assistant Professor Mechanical Engineering Department The University of Texas at Austin

ABSTRACT Recent increases in fuel prices have spurred interest in energy efficiency and alternative energy technologies. These interests are especially relevant for the marine industry, which is responsible for transporting over 90% of the worlds freight. The present global fleet of commercial ships consumes approximately 200 million tonnes of diesel fuel each year which is expected to rise to around 350 million tonnes a year by 2020 [5]. Studies have been conducted evaluating technologies to increase seagoing propulsion efficiency as well as harness available alternative energy sources. One renewable source, wind, is particularly interesting since 1) it presents a vast source of free energy that has been used throughout much of the history of marine transportation, and 2) novel technologies are available that might make it attractive for modern ships. The purpose of this analysis is to specifically evaluate and compare the ability of two modern wind-based technologies to produce thrust-reducing propulsion power for use in reducing the fuel consumption of a ship, namely a rigid wing sail and Flettner rotor. The analysis focuses on design specifications for each based on existing literature and compares the performance of the two technologies within a specified, but naturally varying wind environment. The force-producing capabilities of each technology are compared as a function of the ship operational parameters of heading and speed. INTRODUCTION The dramatic fluctuations of oil prices in the US in the 1970s sparked a lasting interest in oil as an energy supply,

specifically for transportation and as a key component to our economic livelihood. This turmoil spawned significant policy attention to automotive transportation, with an emphasis on fuel efficiency. Similarly, energy efficiency in the marine industry was investigated from a number of perspectives. As with concerns for terrestrial fuel use, once petroleum prices declined back to previous levels many of these perceived issues and investigations were put on the shelf. With concerns about oil prices and global climate change rising, maritime fuel use is gaining attention again. Although total marine fuel use is difficult to define due to its international nature, in 2007 the US consumed 1.5 quadrillion BTU of fuel energy for marine transportation, domestically [1]. According to the most recent projections from the International Energy Outlook in 2009, due to expected growth in international trade, the volume of freight transported by sea is expected to rapidly increase [2, 5]. Thus we are motivated to revive analysis of energy use in the marine industry, and to attempt to develop technological solutions to the challenge of reducing marine fuel use. The intention here is to focus on the design and development of sail technologies for use as propulsion assistance devices, which reduce the load on the conventional propulsion system and its corresponding fuel consumption. Two modern high-lift sail technologiesa rigid wing sail and a Flettner rotorare compared for large-scale merchant ship applications. The technologies are compared in terms of their thrust-producing capabilities for a standard wind profile and ship velocity profile. Before the results are presented, pertinent design criteria are discussed for each mechanism, including

Copyright 2010 by ASME

forward propulsion coefficient optimization. Then, operational details are defined for a simulation, which forms the basis for comparing the two technologies. RIGID WING SAIL DESIGN For over a thousand years, triangular sails have been propelling ships using lift, making them capable of sailing into the wind and producing stronger forces than classic square sails, which are simply pushed by the wind using drag. Thus, the best performing modern sail designs are high lift devices. The most common of these are rigid wing airfoils, such as those used for airplane wings. In accordance with Bernoullis principle, air flows faster over one edge producing a pressure differential. This phenomenon creates a lifting force perpendicular to the direction of airflow as well as a drag force parallel to it. Varying the sail orientation, or incidence angle, manipulates the magnitudes of these lift and drag forces. As illustrated in Figure 1, these forces combine to produce a forward propulsion coefficient, CX, and lateral acceleration coefficient, CY.

Sail geometry must be selected to provide desired lift and drag properties. As mentioned, this type of sail design is fundamentally the same as those used for airplane wings, the primary difference being that here, we desire a symmetric profile to allow lift generation for incidental wind angles from both sides. NACA represents the standard for airfoil design selection and here we particularly focus on the 4-digit series of NACA profiles. These airfoil shapes are defined by 4 single digits, which correlate to values in a standardized shape formula where the first two numbers represent curvature and thus are zero for symmetrical profiles. The last two digits then represent the thickness of the airfoil as a percentage of its chord (length), such that 0015 would represent a symmetrical profile of specific curvature with the maximum thickness being 15% of the total chord length. Comparisons between these different thicknesses are represented in much of the rigid wing sail literature [6, 8]. Increasing the thickness will increase the sail drag; however, this drawback is countered with greater tolerance to rapid small angle variation of wind direction (termed incidence angle). Figure 2 illustrates the maximum lift generated by a profile thickness, represented as a percent of chord length. The result shows a peak at around 18% for profiles such as NACA 0018, with greater values achieved for smoother surfaces [6, 19]. Additionally, when we consider the large amount of mass that must be supported by the mast for these sail systems, substantial concession must be made for fitting the sail mast within the foil profile, providing a structural benefit to the NACA 0018 profile [6, 8].

Figure 1. Lift and drag coefficients for a sail design are resolved to describe forward propulsion and perpendicular forces.

Figure 1 also illustrates that the ship velocity combines with the true wind velocity to create an apparent wind velocity which is the wind speed and direction responsible for force production. The symbol represents the angular difference between ship heading and apparent wind direction, while represents the sail orientation, in degrees, with respect to the incidental wind trajectory. Changes in produce variation in values of CL and CD, which combine to produce CX, as follows: Cx = CL sin CD cos (1)

Figure 2. Maximum lift coefficients for symmetric NACA 4-digit profiles vary as a function of thickness, showing a peak at 18%. [6]

If the device is used as a propulsion assistance device, then the effect of the lateral forces (CY) is minimal and is assumed to be negligible. [11, 12]

Implementing high-lift devices, such as plain trailing edge flaps, has been shown to substantially increase lift generation on the order of over 20%. As complex effects of these devices cannot be modeled accurately [20], data for this investigation were taken from a 30% plain flap with sealed gap and 30 deg maximum angle of deflection was modeled on the trailing edge of the 0018 aerofoil, shown in Figure 3. This design essentially places a hinge for the rear portion (here, 30%) of the wing, with

Copyright 2010 by ASME

the resulting gap sealed by an additional passive device, limited to +/-30 from straight. An additional design factor to consider is the sails aspect ratio the height divided by the chord length, or width. The aspect ratio predominantly affects the rate of change of lift and drag as a function of angle of attack. Generally speaking, high aspect ratios and large sail areas will produce maximum forces, although higher aspect ratios will also produce higher levels of pressure drag. Thus, ships sailing off wind or in low speed conditions benefit from lower aspect ratios and lower sail areas. However, as the aspect ratio decreases towards unity, viscous effects (vortex shedding at airfoil tips) dominate and reduce efficiency. The aspect ratio used here, 5, provides a good balance of all-around performance in varying wind conditions as well as good structural strength, and for this specific sail design was shown to produce favorable lift properties with the lift/drag correlation shown in Figure 3 [8].
2

Figure 4. Illustration of direction of lift produced relative to flow and spin velocity as defined by the Magnus effect concept used in Flettner rotor design. [3]

1.5

Lift Coefficient, Cl

Figure 5. Photograph of one of the original ship prototypes employing a Flettner rotor sail system, c.1924. [http://www.deutsches-museum.de]

0.5

0.2

0.4 0.6 0.8 1 Drag Coefficient, Cd

1.2

1.4

Figure 3. Lift and drag coefficients for a NACA 0018 aerofoil with 30% plain flap, 30 maximum deflection and Reynolds number 0.92 107 [8]

sufficient analytical representation is a problem which continues to be of interest, and development of a closed form solution is still an unsolved problem within the field of fluid mechanics [15]. The textbook representation of the problem consists of the Kutta-Joukowski theorem of lift [3], which utilizes a generalized representation of circulation, commonly symbolized as . This method of quantification produces an analytical approximation of lift, which reduces to . (2)

FLETTNER ROTOR DESIGN SELECTION A Flettner rotor is a sail device that produces lift by spinning a cylinder in an airstream, making use of the Magnus effect. Shown in Figure 4, the Magnus effect describes the fact that when an object spins about an axis perpendicular to airflow, a region of low pressure is formed where the object surface velocity is in the same direction as the flow and a region of high pressure is created where the surface velocity is in the opposite direction of airflow. A lifting force is then exerted on the object towards the low pressure region the same phenomenon behind the motion of a curve ball. Taking the 2 dimensional sketch in Figure 4 as the cross section of a cylinder, as employed in Figure 5, illustrates how this concept can be used as a sail device to produce lift to propel a ship. Calculating the magnitude of the force created by the Magnus effect proves considerably more difficult. The lack of

This equation represents the dimensionless lift coefficient, CL, as a function of the primary operational criterion with respect to operation of a Flettner rotor velocity ratio [3]. Velocity ratio is the dimensionless ratio of tangential velocity of the cylinder surface (v) divided by the free stream velocity (V). As this ratio increases, the lift generated by the Magnus effect generally increases, as shown in Figure 6, where curve (A) illustrates the lift calculated by Equation 2. This linear oversimplification represents what would be the lift produced by a cylinder rotating in an ideal fluid, which is significantly higher than values obtained in practical experimentation. Thus, this value is of little specific use; however, lines (B) and (C) represent 25% and 50% of this approximation, respectively, and it can be seen that the majority of experimental values, (E)(I), commonly lie within this range. Thus, the envelope created from the Kutta-Joukowski theorem produces a design tool, which can be useful in early developmental stages [3]. In

Copyright 2010 by ASME

addition to these inaccuracies, existing analytical representations provide no concession for quantifying drag produced, which must be resolved experimentally. Thus, this experimental directive embodies the most reliable methodology that currently exists for approximating the lift and drag generated by the Magnus effect [3, 4, 6, 7, 8, 9, 11]; and hence this analysis will utilize existing tabulated lift and drag data to calculate propulsion forces from the Flettner rotor. A substantial body of literature was developed in the mid1980s surrounding Flettner rotor applications for ships. In 1986, Borg [3] compiled existing data sets for a wide range of geometries and operating conditions, the relevant subset of which is shown in Figure 6. Each of the different curves shown represents a structural variation in aspect ratio, the inclusion of end plates, different Reynolds numbers, etc, the details of which are shown in Table 1. End plate diameter is expressed as a diameter ratio the diameter of the end plate divided by the diameter of the cylinder.

Much of the analytical literature investigating Flettner rotors or the Magnus effect consists of only lift data, or data in the form of CL/CD with no representation of the relative velocity ratios. Figure 7 represents the relatively small subset of existing data that does include drag measurement. The dynamic analysis undertaken here requires both lift and drag data as a function of velocity ratio. Thus, in conjunction with determining favorable geometry characteristics, our ultimate design selection must be one of the designs represented in Figure 7, such that sufficient experimental data is available.

Figure 7. Experimental data sets for Flettner rotor drag coefficient as a function of velocity ratio [3]

Figure 6. Experimental data sets for Flettner rotor lift coefficient as a function of velocity ratio [3]

Table 1. Geometry specifications for experimental Flettner rotor data [3]

Curve A B C E F G H I L

Aspect Ratio Infinite Infinite Infinite 6.2 13.3 4.7 4.7 4.0 Infinite

End Plate Diameter Ratio None None None 1.58 None 1.7 None 2 None

Reynolds Number Infinite Infinite Infinite 5 4.5x10 1 3.3-11.6x10 4 5.2x10 4 5.2x10 4 11.15x10 4 3.5x10

These charts serve to illustrate many of the key factors involved in Flettner rotor design. One of the most important factors, yet often underrepresented in the literature, is the concept of aspect ratio. This is the ratio of rotor height to rotor diameter. As testing variation of aspect ratio requires construction of an entirely new test apparatus, it is commonly the least varied parameter. Although not well represented by accessible experimental data, it is agreed that a higher aspect ratio produces higher lift given all other criteria the same, and that it significantly reduces drag [3, 6], although these factors must be considered in conjunction with structural requirements. Few explicit rules of thumb exist, but an optimum value of 6 has been proposed [3], while it has been concluded that ratios below 4 are unfavorable [11]. Data sets (G) and (H), in Figure 6 and 7, represent identical configurations with the exception that (G) has been fitted with end plates. The use of end plates is a common practice, of obvious benefit, consisting of placing flat discs of diameter slightly larger than the cylinder itself on the end of the rotor.

Copyright 2010 by ASME

These components also increase the required drive torque with increasing diameter, and the suggested practical range is of 1.5 to 2 times the base cylinder diameters. These effects have been studied in significant detail to show that there is potential benefit of lift increase by introducing additional discs equally spaced along the cylinder [18], although practical utilization has yet to prove this benefit and there is likely a very large increase in driving torque demand. Flettner rotor sail designs also provide significant concession for safety during dangerous storm or high-speed wind conditions [3, 6]. This benefit of the Flettner rotor lies in the behavior at velocity ratios around 1 when cylinder surface velocity is equivalent to the airflow. At this point the drag coefficient, as shown in Figure 7, drops below levels experienced at stand still. Known as the Barkley phenomenon, it gives the Flettner rotor tremendous safety advantage because the Flettner rotor is able to reduce its presence to below that of a full rigged vessel operating with bare poles [3, 6]. It is also important to note that it was shown that within this range of velocity ratio less than unity was when lift forces experienced substantial fluctuation as a function of Reynolds number meaning that lift varied as a function of absolute speed in addition to speed ratio. Although theoretically some limit should exist, within the range studied here, Reynolds number is assumed to be irrelevant [16]. These geometric ideals focus our options to data sets (E), (I) and (L). (L) represents experimental data from a test set up using an infinite aspect ratio, which thus lends insight into the nature of drag development being a function of effects at cylinder end. Designs (E) and (I) both incorporate end plates (of 1.58 and 2 times the cylinder diameter, respectively). Configuration (E) provides enhanced performance as shown by its higher lift coefficients (at the cost of increased drag) due to its higher aspect ratio; however, this study will utilize (I) as it meets identified design criteria and has complete data through a velocity ratio of zero. Of final consideration with regard to the Flettner rotor is the need to approximate the required input torque. If the system is to operate at a variable speed in response to wind velocity fluctuation a significant amount of energy must be put into driving this system. Only a few investigations have ventured to include analytical representation of this factor and all of these methods are derived from surface friction approximations. The most practical evaluation is based on an evaluation of the coefficient of friction using Reynolds number such that (3) Where crot is the velocity ratio, LRe is the characteristic length (circumference of the rotor), is the dynamic viscosity of air and A and VA are the density and speed of airflow. This is then used to calculate the frictional coefficient [4], (4) such that frictional force can be calculated as

(5) from Ar, the surface area of the rotor and Urot, the surface velocity. From this the torque can be calculated by multiplying Ff by radius, or power by multiplying Ff by cylinder rotational velocity. In addition to this analytical method, some studies simply consider input power to be a percentage of generated power. Providing the most substantiated of assumed drive power requirements, Bergeson [6] refers to model-calculated results of 50-66% fuel savings from integrating the Flettner sail system. Experimental results revealed 20% savings when fuel used to drive the motor was accounted for, suggesting that 3046% of the energy saved by the wind-assist system was used to drive the rotor. Due to the large uncertainty associated with all of these figures, as well as the importance of this concept, each of these valuations is assessed in the final analysis. SIMULATION: OPERATIONAL PARAMETERS The previously selected designs for sail and rotor determine the values of lift and drag characteristics for the configuration. Given these specifications, we can run an optimization routine to determine the maximum potential forward propulsion coefficient, CX (referring to Figure 1), for the entire range of potential incident wind angles. To conduct this optimization, for each apparent wind angle, , we vary the sail angle (for the wing sail) and velocity ratio (for the Flettner rotor) over the entire feasible range to determine what resulting combination of CL and CD produces the maximum CX, as in Equation 1. The apparent wind angle range is restricted to 0180 as these results are symmetric (an apparent wind angle of 90 will produce the same CX as an angle of 270). The results of this optimization are shown in Figure 8, where 0 indicates sailing directly into the wind, where we can produce no forward force, and 180 corresponds to sailing directly with the wind, where a nominal amount of drag allows us to still produce some force.
15 Flettner Rotor Wing Sail

10

Cx
5 0 0

20

40

60 80 100 120 Incident Angle (deg)

140

160

180

Figure 8. Forward propulsion coefficient (CX) results from optimizing control parameters for each sail design, where 0 represents sailing directly into wind and 180 sailing parallel with wind.

Copyright 2010 by ASME

A wind profile of fairly even speed distribution (mean speed of 5.3 m/s, peak of 11.2 m/s) is taken from a datacollecting buoy in the Atlantic Ocean, along the common trade route between the US and Britain [14]. Wind speed and direction averages are given hourly for an 11 day time period in May 2008. This profile is selected as it displays wind directions predominantly of a single direction, with some noise as desired for representing natural fluctuation, with a histogram of the direction distribution shown in Figure 9. Here we can see that wind blows primarily from 180, meaning directly from the south. Thus, in the actual simulation, the majority of the wind is approaching the notional ship from the rear and hence a heading of 180 will represent sailing most directly into the wind. Results will be based on variations of that heading direction.

continuous one following the ships trajectory again reflecting more of a standardized comparison of sail technologies than explicit performance for a trans-Atlantic voyage. The power generated is calculated from Equation 6. (6) To this point all pertinent geometries and velocities required for calculating sail power as in Equation 6 have been analyzed except for the sail area. The relative scales used by Wind Ship Company [6], derived from experience, were used for these analyses: 3,000 ft2 wing sail and 360 ft2 rotor. Then, a simulation was executed for a defined ship heading relative to the wind profile. The power required to drive the Flettner rotor is then subtracted from the propulsive power it produces, calculated both analytically and using the given experimental metrics. The result is a profile of power produced by each system as a function of time, such as the example illustrated in Figure 10. As shown, each system undergoes periods of minimal power production, due to unfavorable wind directions or speeds, and periods of higher performance. The average power produced over the length of the simulation by each unit (net for the rotor) is used to illustrate each technologys performance comparable to the other systems for that specific ship heading. The results of such a comparison for the ship speeds indicated earlier (6, 8, 10 and 12 m/s) are shown in Figures 11-14. Here the range between the high and low experimental estimates for losses from drive torque input is represented by the shaded region.

300

Figure 9. A histogram of the wind direction distribution for the profile used for simulation shows a concentration at 180

250

Rotor - Analytical Losses Rotor - Experimental Losses (HI) Rotor - Experimental Losses (LO) Rigid Wing Sail

This framework thus assumes that each device is able to instantaneously adjust to current wind conditions, which is valid from a propulsion perspective as wind data is given as hourly average. However, this assumption does not take into account the energy consumed in accelerating the rotor nor repositioning the sail, which are assumed to be of similar enough magnitude to cancel each other out. SIMULATION: COMPARISON AND RESULTS This analysis approaches the comparison from the standpoint of quantifying the propulsion power produced by a single sail of each design fitted to a hypothetical ship. This hypothetical ship has a speed and direction which affects the performance of the sail, but the ability of each sail mechanism to produce useful propulsion forces is independent of any other ship parameter. Ship speed is assumed to be constant over the 11 day period, and performance is evaluated at speeds of 6, 8, 10 and 12 m/s. As the actual wind data used are for a single stationary spot it is thus assumed that this profile would be a

200

Power (kW)

150

100

50

0 80

100

120

140 Time (hr)

160

180

200

Figure 10. Instantaneous propulsion power produced by each design for a sample time at a heading of 40 and a speed of 8m/s (HI represents the high estimate of experimental rotor losses, LO indicates the lower estimate)

Copyright 2010 by ASME

Figure 11. Average power produced by each sail model for a notional ship speed of 6 m/s as a function of heading.

Figure 13. Average power produced by each sail model for a notional ship speed of 10 m/s as a function of heading.

Figure 12. Average power produced by each sail model for a notional ship speed of 8 m/s as a function of heading.

Figure 14. Average power produced by each sail model for a notional ship speed of 12 m/s as a function of heading.

Generally speaking, Figures 11-14 show that the selected sail and rotor sizes, or areas, used for comparison provide a strong, even basis for contrasting the capabilities of each. Depending on vessel speed, once ship heading approaches the range of 30-60, it can generally be said that wing sail performance is easily within the range of experimental uncertainty for Flettner rotor power production. As ship speed increases, these two regions become similar over an increasing range, to the point where at a velocity of 12 m/s, they are the same for the majority of the spectrum, with the wing sail actually producing increased performance for headings most directly into the wind. It is also clear that for scenarios where the ship heading is in the opposite direction of incident wind, which corresponds to a low heading (ie 0), the Flettner rotor is able to produce better

performance, a logical conclusion from its significantly higher drag values. The fact that the Flettner rotor produces proportionately higher drag along with its increased lift also indicates that applications where lateral accelerations are of concern (smaller ships, or high speed ships) would have less benefit from the Flettner rotor. For all scenarios, the analytical approximation of the amount of torque and power required to drive the Flettner rotor indicates significantly lower power demands, and thus higher net power levels of production. The methodology used to derive this number is based on a simplified model, thus additional verification is warranted. Conversely, losses due to drive torque as derived from experimental data are based on measured fuel quantities. Thus, these figures likely take into account an efficiency loss through the combustion of the

Copyright 2010 by ASME

engine, where as this analysis is chiefly comparing useful, mechanical power. Therefore, from an energy or power perspective, the actual amount of energy used to rotate the cylinder is less than the total amount of energy consumed in the form of fuel, and losses from this are potentially smaller than indicated. DISCUSSION AND CONCLUSIONS The primary conclusion of this analysis is that for a cross sectional area of almost an order of magnitude less, a Flettner rotor is able to produce equivalent if not better propulsion performance when compared to the best practice wing sail design, including all reasonable estimates for required drive torque. In addition, if we consider this analysis within the scope of retrofitting to large-scale merchant ships, such as container and cargo ships, deck space utility is a primary concern. Inhibiting the capacity to store cargo, as well as restrictions to the use of cranes to load and unload cargo are vital to the economic livelihood of the ships. The factors at play are both deck area occupied as well as system height. A wing sail design will require substantially more area to operate as it must be able to freely rotate 360. This requirement leads to a circular footprint with a radius of up to the sail chord length. In contrast, the Flettner rotors footprint is the circular cross sectional area of the cylinder, which does not vary with time. Additionally, as can be seen in the figure below, the smaller area of the Flettner rotor compared with the rigid sail produces a significantly reduced height. These factors combine to indicate that the Flettner rotor design produces a sail system weighing only 22% of the equivalent sail system [6].

speed ships, or ships which intend to derive the majority of their propulsion from wind power, the high drag that accompanies the Flettner rotors high lift may be of specific concern. Absolute conclusions on the capabilities of each design rely on the accuracy and methodology for calculating Flettner rotor drive torque. Experimental evidence has a significant range of uncertainty and has a substantial discrepancy with analytical approximations. The analytical approximations used here are based on surface friction factors on the cylinder, which also represents a large degree of uncertainty. Additionally, as the surface/air interactions produced by the Magnus effect are not well understood, it is likely that these same effects would impact drive torque requirements; particularly given a representation utilizing surface friction for calculation. These observations suggest the need to collect data regarding drive requirements in conjunction with the ongoing need for more lift and drag force data. Despite these uncertainties, for the application studied here of providing sail-assist performance in the interest of reducing fuel consumption on large scale ships while taking up minimal deck space, we see that the Flettner rotor, although needing further testing and verification, certainly presents a comparable if not superior design alternative to the wing sail. REFERENCES [1] Davis, S., Diegel, S., 2009, "Transportation Energy Data Book: Edition 28," Oak Ridge National Lab, Center for Transportation Analysis, ORNL-6984, Oak Ridge, TN. [2] International Energy Outlook, 2009. U.S. Energy Information Administration, Report # DOE/EIA-0484. [3] Borg, J., 1986, "The Magnus Effect - An Overview of its Past and Future Practical Applications," The Borg/Luther Group, Naval Sea Systems Command Contract #N00024-83-C-5350 [4] Silvanius, M., 2009, "Wind Assisted Propulsion for Pure Car and Truck Carriers," Master's thesis, KTH Centre for Naval Architecture, Stockholm, Sweden. [5] Shukla, P., Ghosh, K., 2009, "Revival of the Modern Wing Sails for the Propulsion of Commercial Ships," Int. J. of Environmental Science and Engineering, 1(2), pp. 75-80. [6] Bergeson, L., Greenwald, C., 1985, "Sail Assist Developments 1979-1985," Proc. International Symposium on Windship Technology 4A (WindTech '85), Southhampton, U.K. [7] Smulders, F., 1985, Exposition of Calculation Methods to Analyse Wind-Propulsion on Cargo Ships, Proc. International Symposium on Windship Technology 4A (WindTech '85), Southhampton, U.K. [8] Fiorentino, F., 1985, Exposition of Calculation Methods to Analyse Wind-Propulsion on Cargo Ships, Proc. International Symposium on Windship Technology 4A (WindTech '85), Southhampton, U.K.

Figure 15. Illustration of the relative scale of each sail system, with the 3,000 ft2 wing sail on top and 360 ft2 rotor on bottom [6].

The different designs also have different performance characteristics. The rotor provides significantly better performance at low speeds and under directions following closely with the wind. As ship speed increases, the performance of each design begins to normalize, indicating potential benefit for the rigid wing sail. Additionally, for high

Copyright 2010 by ASME

[9]

[10]

[11]

[12]

[13]

[14]

[15]

[16]

[17]

[18]

[19]

[20]

Shenzle, P., 1985 "Estimation of Wind Assistance Potential," Proc. International Symposium on Windship Technology 4B (WindTech '85), Southhampton, U.K. Satchwell, C., 1985, "Aerodynamic Design of Aerofoils," Proc. International Symposium on Windship Technology 4B (WindTech '85), Southhampton, U.K. Wellicome, J., 1985, "Some Comments on the relative merits of Various Wind Propulsion Devices," Proc. International Symposium on Windship Technology 4B (WindTech '85), Southhampton, U.K. Wilson, P., 1985, "A Review of the Methods of Calculation of Added Resistance for Ships in a Seaway," Proc. International Symposium on Windship Technology 4B (WindTech '85), Southhampton, U.K. Ingham, P., Terslov, O., 1985, "Wind Tunnel Tests and Manoeuvre Simulator Tests with Different Types of Sails and Ships," Proc. International Symposium on Windship Technology 4B (WindTech '85), Southhampton, U.K. National Data Buoy Center, 2008, Station 41001, LLNR 635, Historical Climate Data, http://www.ndbc.noaa.gov/station_page.php?station=410 01. Badr, H., Coutanceau, M., Dennis, S., Menard, C., 1990, "Unsteady Flow Past a Rotating Circular Cylinder at Reynolds numbers 103 and 104," J. Fluid Mech., Vol. 220, pp. 459-484. Swanson, W., 1961, "The Magnus Effect: A Summary of Investigations to Date," Transactions of the ASME, Journal of Basic Engineering, Paper no. 60-WA-150, pp 461-470. Fujiwara, T., Hearn, G., Kitamura, F., Ueno, M., 2005, "Sail-Sail and Sail-Hull Interaction Effects of HybridSail Assisted Bulk Carrier" J. Marine Science Technology, V10, pp 82-95. Clayton, B., 1985, "BWEA Initiative on Wind Assisted Ship Propulsion (WASP)," Proc. International Symposium on Windship Technology 4A (WindTech '85), Southhampton, U.K. Goett, H., Bullivant, K., 1939 "Tests of the NACA 0009, 0012, and 0018 Airfoils in the Full-Scale Tunnel," NACA Report 647. Abbott, I., Von Doenhoff, A., 1959, Theory of Wing Sections, Including a Summary of Airfoil Data, Dover Publications, Inc., New York, Chapter 8.

Copyright 2010 by ASME

Potrebbero piacerti anche