You are on page 1of 23

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/258071312

Summary of Savonius wind turbine development and future applications for


small-scale power generation

Article in Journal of Renewable and Sustainable Energy · August 2012


DOI: 10.1063/1.4747822

CITATIONS READS

84 15,494

5 authors, including:

John Patrick Abraham B. D. Plourde


University of St. Thomas University of St. Thomas
308 PUBLICATIONS 5,623 CITATIONS 39 PUBLICATIONS 550 CITATIONS

SEE PROFILE SEE PROFILE

Greg Mowry E. M. Sparrow


University of St. Thomas University of Minnesota Twin Cities
43 PUBLICATIONS 747 CITATIONS 842 PUBLICATIONS 31,083 CITATIONS

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Numerical modelling of hydraulic structures View project

Screens on the hydraulic performance of a drops View project

All content following this page was uploaded by John Patrick Abraham on 26 August 2014.

The user has requested enhancement of the downloaded file.


Summary of Savonius wind turbine development and future applications for
small-scale power generation
J. P. Abraham, B. D. Plourde, G. S. Mowry, W. J. Minkowycz, and E. M. Sparrow

Citation: J. Renewable Sustainable Energy 4, 042703 (2012); doi: 10.1063/1.4747822


View online: http://dx.doi.org/10.1063/1.4747822
View Table of Contents: http://jrse.aip.org/resource/1/JRSEBH/v4/i4
Published by the American Institute of Physics.

Related Articles
Control strategies of doubly fed induction generator-based wind turbine system with new rotor current protection
topology
J. Renewable Sustainable Energy 4, 043123 (2012)
Passive load alleviation bi-stable morphing concept
AIP Advances 2, 032118 (2012)
Correlation analysis for wind speed and failure rate of wind turbines using time series approach
J. Renewable Sustainable Energy 4, 032301 (2012)
Electromagnetic design analysis and performance improvement of axial field permanent magnet generator for
small wind turbine
J. Appl. Phys. 111, 07E708 (2012)
Review: The use of geographic information systems in wind turbine and wind energy research
J. Renewable Sustainable Energy 4, 012701 (2012)

Additional information on J. Renewable Sustainable Energy


Journal Homepage: http://jrse.aip.org/
Journal Information: http://jrse.aip.org/about/about_the_journal
Top downloads: http://jrse.aip.org/features/most_downloaded
Information for Authors: http://jrse.aip.org/authors
JOURNAL OF RENEWABLE AND SUSTAINABLE ENERGY 4, 042703 (2012)

Summary of Savonius wind turbine development and future


applications for small-scale power generation
J. P. Abraham,1,a) B. D. Plourde,1,b) G. S. Mowry,1,c) W. J. Minkowycz,2,d)
and E. M. Sparrow3,e)
1
School of Engineering, University of St. Thomas, 2115 Summit Ave, St. Paul,
Minnesota 55105-1079, USA
2
Department of Mechanical and Industrial Engineering, University of Illinois, Chicago,
842 Taylor St., Chicago, Illinois 60607, USA
3
Department of Mechanical Engineering, University of Minnesota, 111 Church St,
Minneapolis, Minnesota 55455-0111, USA
(Received 15 May 2012; accepted 10 August 2012; published online 23 August 2012)

Wind turbine use is expanding throughout the world as a means to provide


electricity without contributing to the increase in global-warming gases. Most
commonly, very large, horizontal-axis turbines are constructed in fleets that are
connected to national-level electrical grid systems. More recently, there has been a
desire for more local, small-scale power production that can be used to power very
specific pieces of equipment or buildings. Some of the small-scale turbines are
designed differently from their larger counterparts—they are driven by drag forces
rather than by lift. Drag-driven turbines are typically called Savonius turbines. This
paper, which presents a historical perspective on Savonius turbines, will illustrate
their potential for providing local power. Finally, we will discuss recent
developments in analysis methods which intend to optimize Savonius turbines for
powering cellular communication towers in developing parts of the world. V C 2012

American Institute of Physics. [http://dx.doi.org/10.1063/1.4747822]

INTRODUCTION
Wind power is one of the most promising sources of future energy for the world. Global
energy availability is far in excess of the energy required to power the globe’s current electrical
usage, even though current technologies only extract a small percentage of the available wind
power.
Most wind power is generated from very large wind turbines that are constructed in large
wind farms with dozens or hundreds of turbines that are connected to regional or national elec-
trical grids. Current worldwide capacity is approximately 240 000 MW (Ref. 1) and is growing
significantly each year.
Concurrent with large-scale grid-level wind-power production, which has now expanded to
offshore turbines whose sizes exceed those which can be used on land, a new focus has been
given to small-scale local wind power production. For this type of application, the power is to
be produced at the site of use. Because of the nature of local production and utilization, small-
scale systems benefit from the use of power storage capacity for situations in which the power
generated exceeds that required for the application. Later, the stored power can be released

a)
Author to whom correspondence should be addressed. Electronic mail: jpabraham@stthomas.edu. Tel.: þ1-651-962-
5766; Fax: þ1-651-962-6419.
b)
E-mail: bdplourde@stthomas.edu.
c)
E-mail: gsmowry@stthomas.edu.
d)
E-mail: wjm@uic.edu.
e)
E-mail: esparrow@umn.edu.

1941-7012/2012/4(4)/042703/21/$30.00 4, 042703-1 C 2012 American Institute of Physics


V
042703-2 Abraham et al. J. Renewable Sustainable Energy 4, 042703 (2012)

when the reverse situation occurs. Energy storage is typically carried out through lead-acid bat-
teries or raised water; both of which are readily available around the world.
Whether wind power is used on a small or large scale, a number of design considerations
have to be given prior to its introduction. First, wind power is inherently intermittent; energy
storage or auxiliary power sources must be employed. It is common, for instance, to find small-
scale wind and solar systems linked together into a single unit. Alternatively, grid connectivity
or diesel generation can provide a reliable source of backup power when wind speeds are not
sufficient to generate electricity.
For small-scale power production, which is defined here to be on the order of a few kW,
large horizontal-axis wind turbines (HAWTs) are not appropriate. A different class of turbines
is often used which relies upon rotation about a vertical axis and are therefore classified as
vertical-axis wind turbines (VAWTs).
VAWTs have two major variants, namely, Savonius and Darrieus. While Savonius turbines
are the focus of this work, a discussion of the distinction between the two is warranted. While
both have vertical axes about which they rotate, the aerodynamics are very different. Darrieus
rotors appear as long airfoil structures that cause rotation by lift forces. Darrieus turbines are, at
least with respect to lift, similar to HAWTs.
Savonius rotors are driven by wind drag forces. Consequently, they have a very different
appearance compared to their lift-driven counterparts. In their most rudimentary design, a Savo-
nius rotor is created by bisecting a cylinder and offsetting the cylinder halves about a central
axis of rotation. Figure 1 has been prepared to show a schematic of a typical Savonius rotor
with the most important dimensions listed in annotations in the figure.
As indicated in the figure, the major parts of a traditional Savonius rotor are the radius of
the rotor and the degree of offset (or overlap) of the two halves. There also may be a separation
distance, as indicated. Since the image of Fig. 1 is a two-dimensional profile, it does not show
the height of the turbine—which would be a third critical dimension.
In some applications, the turbines have half-caps at their upper and lower surfaces, particu-
larly if they were manufactured from bisected oil drums. An image of an obliquely oriented
Savonius turbine showing a cap is provided in Fig. 2.
The dramatic difference in design between HAWTs, Darrieus rotors, and Savonius rotors is
reflected in their vastly different performance. The performance of a turbine is often measured
by the value of its power coefficient. This term is, in essence, an efficiency. It is a measure of
the extracted power compared to the power of wind flowing through the turbine’s projected
area. Mathematically, the power coefficient (Cp) is expressed as

FIG. 1. Top-down view of a Savonius-style rotor.


042703-3 Abraham et al. J. Renewable Sustainable Energy 4, 042703 (2012)

FIG. 2. An oblique view of a capped Savonious rotor.

Power
Cp ¼ 1 3
: (1)
2 AqV

Here, Power is the power extracted from the turbine, A is the rotor projected area, q is the air
density, and V is the wind speed. Through very rudimentary fluid-mechanic analyses, it is
shown that the maximum possible Cp is 59.3%. This limit is termed the Betz Limit.
Modern, large HAWTs can achieve Cp values up to 0.5 (50%). VAWTs however, particu-
larly Savonius rotors, have much lower performance, approximately 15%. Because of this, it is
legitimate to inquire whether there are any advantages that VAWTs can offer to overcome this
performance deficit.
Among the critical differences between Savonius VAWTs and HAWTs are the following:
(1) VAWTs occupy a smaller footprint so they can be used in confined physical locations or
can be positioned close together;2 (2) VAWTs do not need control systems to point the rotors
into the wind, they are able to rotate regardless of wind direction; (3) Savonius rotors are able
to start rotating with slower wind speeds and can, consequently, generate power for these low
wind speeds; and (4) Savonius turbines rotate slower than lift-driven rotors so they impart a dif-
ferent mechanical load to their support structure. Because of these differences, in certain situa-
tions, a VAWT turbine is the preferable option.
One instance in which a VAWT is advantageous is when a turbine is attached to a
building and is used to provide local electricity. Often, buildings have large equipment such
as heating, ventilation, and air-conditioning (HVAC) units which create obstructions to air
flow or to the positioning of the turbine. A VAWT can be placed in more confined spaces
than a HAWT.
Other applications include the use of VAWTs to provide local electricity for residents of
remote villages in the developing world. Small, robust VAWT systems with battery storage
capability enable the intermittent generation and storage of power which can be used for small
electronics, charging of cellular phones, or lighting.
042703-4 Abraham et al. J. Renewable Sustainable Energy 4, 042703 (2012)

Another very promising application for VAWTs is to power cellular communication tow-
ers. Around the world, growth in cellular communication is strong with many countries bypass-
ing more traditional land-line phone networks. A cellular communication network requires
adequate coverage for its users. Consequently, a sufficient number of communication towers
with antennas are needed for this coverage. The typical power requirements are 1-3 kW for
cell-phone electronics associated with a modern tower. In some instances, this energy is pro-
vided by grid-based electricity. For many parts of the world where grid electricity is not contin-
uous, power outages result in a collapse of communications. Such outages are often prevented
through the use of diesel backup generation (generator-motor systems). In other instances, the
power grid is not connected to the tower. This is particularly the case in very rural areas. In
these situations, a local source of power is critical to the operation of the network.
Figure 3 has been prepared to show a vertical-axis turbine in the aforementioned applica-
tion (powering cellular communication towers). The image shows that in a preferred applica-
tion, the turbine is installed well above local obstructions such as treelines so that windspeeds

FIG. 3. Savonius turbine attached to a cellular communication tower. Reprinted with permission from J. P. Abraham, B. D.
Plourde, G. S. Mowry, E. M. Sparrow, and W. J. Minkowycz, J. Renewable Sustainable Energy 3, 033109 (2011). Copy-
right 2011 American Institute of Physics.
042703-5 Abraham et al. J. Renewable Sustainable Energy 4, 042703 (2012)

are maximized. An obvious benefit of the intended application is that the turbine installation
does not require the construction of a separate tower. The illustration of Fig. 3 is meant to be
illustrative of typical wind variations and is intended to be suggestive of the installation. The
turbine drawn in that figure is not to scale.
A series of photographs have been prepared to show the actual installation of a turbine,
such as that shown in Fig. 4.5
A close-up view of a turbine attached to the tower is provided in Fig. 5. There, a three-
sectioned rotor blade is shown with the supporting structures above and below the turbine. In
addition, there is a generator positioned immediately below the rotor.
It is common to refer to the small-scale wind systems described here as distributed systems
whether they are lift-driven or drag-driven. A workshop in 2008 (Ref. 5) came to several clear
conclusions on the future of distributed systems. Among them: (1) continued research is needed
on airfoil design, gearboxes, stall control, among other areas; (2) distributed systems are a key
component to a goal of achieving significant wind power penetration into the power production
market; (3) community acceptance is an important component to the success of distributed

FIG. 4. A Savonius turbine attached to a communication tower. Reproduced with permission from B. D. Plourde, J. P.
Abraham, G. S. Mowry, and W. J. Minkowycz, Sens. Transducers J. 13, 53-61 (2011). Copyright 2011 International
Frequency Sensor Association.
042703-6 Abraham et al. J. Renewable Sustainable Energy 4, 042703 (2012)

FIG. 5. Close-up photograph showing the supporting structures which connect the turbine to the tower and the generator
situated immediately beneath the turbine. Reprinted with permission from J. P. Abraham, B. D. Plourde, G. S. Mowry,
E. M. Sparrow, and W. J. Minkowycz, J. Renewable Sustainable Energy 3, 033109 (2011). Copyright 2011 American
Institute of Physics.

wind production; (4) energy storage improvements will be critical to distributed systems; and
(5) improvements are needed in the efficiency, control, and cost of the power electronics.

TECHNICAL BACKGROUND
There have been a large number of research studies on Savonius-style turbines. The litera-
ture survey set forth in the following is intended to show the chronological development of
Savonius rotor development. While it is not intended to be completely exhaustive, it does cap-
ture the principle works in this field.
Prior to the discussion of literature, it is important to review the basic operation of a Savo-
nius rotor which is best displayed by reference to Fig. 6. The figure shows a cross-sectional cut
through a Savonius-style rotor. The rotor is shown to spin in a clockwise direction while airflow
patterns are shown by streamlines which were extracted from prior two-dimensional simula-
tions. The streamlines are directed from the bottom of the image upward. There are two torques
which are imparted on the turbine. The primary torque promotes rotation in the direction of
rotation and an opposing torque serves to retard (slow) the rotation of the device. In order to
optimize performance, the primary torque should be maximized while the opposing torque
should be minimized.
The primary mechanism for optimizing turbine performance is through field-testing or
wind-tunnel testing. After the initial introduction of drag-style rotors was made in Ref. 6, it
was a number of years before serious investigations followed. Among the first studies which
042703-7 Abraham et al. J. Renewable Sustainable Energy 4, 042703 (2012)

FIG. 6. Streamlines and the primary and opposing torques which cause rotation of a Savonius rotor. Reprinted with permis-
sion from J. P. Abraham, B. D. Plourde, G. S. Mowry, E. M. Sparrow, and W. J. Minkowycz, J. Renewable Sustainable
Energy 3, 033109 (2011). Copyright 2011 American Institute of Physics.

experimentally investigated these rotors were Refs. 7 and 8 which evaluated the importance of
tunnel blockage on the drag forces incurred on a Savonius rotor. The second of those works
investigated not only the drag forces but also the efficiency with which the rotor can extract
energy. They reported results for a variety of different configurations. All results were obtained
with scaled-rotors positioned in a wind tunnel. It was found that a standard Savonius turbine
could reach efficiency levels of 0.0653. However, with modifications such as capping, efficien-
cies up to 0.25 were achievable. In that work, the efficiency (g) was related to the so-called
Betz Limit efficiency of 0.593

Power 1
g¼1 3 0:593
: (2)
2 AqV

Alexander and Holownia8 also used the tip-speed ratio (k) to show how performance depends
on the rotational speed of the device. The definition of k is

Vtip
k¼ : (3)
Vwind

They showed consistently that performance peaks for k values in the range from 0.4 to 0.7 and
that performance dropped off significantly for values of k either above or below this range.
Also in 1978, an article was published which investigated a large number of Savonius-style
rotor designs.9 They too dealt with the effect of wind-tunnel blockage as they experimented on
two- and three-bucket Savonius rotors. Their experiments covered a range of Reynolds num-
bers, bucket configurations, and bucket overlaps. They concluded that two rotor sections, each
with a slight overlap and positioned with a 90 angular offset provided maximum performance.
The authors also concluded that it is possible to achieve power coefficients that range from
approximately 0.14 to 0.25 with peak values occurring for tip-speed ratio values of approxi-
mately 0.8. These values are a bit higher than prior reporting of performance, partly because
the authors included frictional torque losses in their results.
042703-8 Abraham et al. J. Renewable Sustainable Energy 4, 042703 (2012)

In the next decade, a series of studies repeated the works of Refs. 6–8 and came to similar
results. For instance, in Ref. 10, oil drums were bisected to create the Savonius rotor. The
rotors were tested with various overlap conditions. It was found that power coefficients on the
order of 0.06–0.08 could be achieved at corresponding values of k that were in the 0.3–0.5
range. It should be noted that Ref. 10 used the term “performance coefficient” to refer to what
is here called the “power coefficient.” It was also shown in this work that a negative offset
allowed improvements in performance; this finding had not been previously reported in the lit-
erature. Another conclusion was that the use of end plates significantly increased performance
and raised the tip-speed ratio at which optimal performance occurred.
Shortly thereafter, a very detailed study on the geometric parameters governing Savonius
turbines was performed in Ref. 11. In this study, the scaled turbine was placed just downstream
of a wind-tunnel exit; no blockage effects were taken into consideration. The study evaluated
seven separate issues: (a) rotor aspect ratio, (b) overlap, (c) separation between buckets, (d) pro-
file of the buckets, (e) number of buckets, (f) presence or absence of endplates, and (g) effect
of multiple stages of rotors. The investigation also included flow visualization studies to learn
about the patterns of fluid flow around the blades. It was found that very high power coeffi-
cients could be achieved (approximately 0.25) and that performance fell steeply as the tip-speed
ratio changed. The results of Ref. 11 are difficult to rationalize with those of prior works which
showed significantly lower performance. Part of the difference could be the details of the
experiments and the inclusion or exclusion of blockage effects.
Adding to the literature, Ref. 12 reported performances of Savonius style turbines that
were relatively high compared to prior works. The discussion in Ref. 12 pointed out that prior
studies were made with small wind tunnels and blockage effects were either ignored or under-
appreciated. Among the considerations in Ref. 12 were bearing friction loss, blockage, multiple
turbine sections, as well as many different geometrical parameters which define the rotor. The
investigators also measured local pressures using rotor wall pressure taps for non-rotating devi-
ces and they proposed a modified quasi-steady approach which can relate steady state flows to
those flows which occur during actual turbine rotation. In conclusion, the authors report a peak
power coefficient of 0.32 which corresponds to a tip-speed ratio of 0.79. The dimensions for
this optimal device were provided in great detail in this paper.
Also in 1989, a study was published13 that investigated the impact of the use of deflecting
plates which were positioned forward of the Savonius rotors that found optimized plates could
increase performance substantially (30%) compared to the no-plate case.
Whereas most of the aforementioned studies were focused on optimizing performance of
the Savonius rotor, a few studies attempted to make detailed investigations of the flow patterns
and pressure distributions which occur in the fluid itself. For instance, Ref. 14 compared both
still and rotating rotors. They confirmed expectations that negative pressures on the backside of
the returning blade presents a detriment to the overall performance of the device. On the other
hand, they noted significant differences in the values of pressure for still and rotating devices.
A continuation of the study by the same authors, published two years later15 provided essen-
tially the same conclusions as those set forth in Ref. 14.
Shortly thereafter, two theses were published which dealt with experimental and computa-
tional investigations of Savonius-style rotors that included blade twisting and valves.16,17 Both
of these works presented updated discussions on the state-of-the-art at the time. Of special note
is that in Ref. 17, a two-dimensional simulation was performed, however, the flow was treated
as laminar (Reynolds number based on diameter 100 000) which is unlikely, particularly
downstream of the object. First, shed vortices tend to have a three-dimensional structure. Also
of note is that turbulent flow may be established by natural transition where small disturbances
grow in the streamwise direction; by bypass transition, where the flow may skip natural transi-
tion and pass directly to vortex breakdown; or by separated-flow transition wherein a laminar
boundary layer separates.18 The factors which are present here that promote bypass or
separated-flow transition include the following: (1) freestream turbulence—even very small val-
ues of freestream turbulence significantly impacts the transition,18 (2) adverse pressure gra-
dients, (3) counter-moving surfaces and fluids, (4) edges that may serve as a boundary-layer
042703-9 Abraham et al. J. Renewable Sustainable Energy 4, 042703 (2012)

tripwire. Recent studies that have dealt with simulations of laminar-to-turbulent transition are
Refs. 19–30.
Additionally, in Ref. 17, the modeling effort did not seem to include full rotation motion.
In the text, a static case was described as well as two orientations for which the rotors were
provided a rotational speed. Consequently, the impacts of changing flow patterns during a full
rotation cycle could not be dealt with.17 The works of these theses appeared in the follow-on
published report of Ref. 31. Another output from the same research group32 investigated valve-
assistance in improving the performance of Savonius blades which suggested that one-way
valves may be able to improve the performance of large-scale Savonius rotors. At least for
small-scale Savonius rotors, the valve-assisted designs in Ref. 32 did not appreciably affect per-
formance and were prone to vibration and noise problems.
Later, in Ref. 33, an optimization study was performed on the impact of blade twisting
of Savonius rotors which showed a marked improvement in the power coefficient with angle
of rotation (up to 15 ). Even more recently,34 that research group carried out a series of
wind-tunnel experiments on Savonius rotors with various stages, valving and capping to
optimize the design. They tested rotors with and without twists. They found extremely high
values of the power coefficient, sometimes in excess of 0.3. To the best knowledge of the
authors, the results found in Ref. 34 are among the highest reported in the literature. If these
results could be confirmed, it would represent an extraordinary potential for Savonius-
generated power.
With respect to a summary of Savonius performance, it has typically been found that
power coefficients on the order of 0.15 are possible, in some instances have results exceeded
this level. It has also been found that the power coefficient is strongly dependent on the tip-
speed ratio. As the tip-speed ratio moves from its optimum value, the performance drops signif-
icantly. There is a strong variation of performance reported in the literature and this variance
leads to some question on the potential of Savonius turbines. It is the opinion of the present
authors that much of the variation can be attributed to experimental technique, the accounting
for or neglect of frictional resistance, losses in converting mechanical energy to electrical
power, and blockage corrections.
Very recently, the present authors have begun a series of studies which include both the
simulation and experimental validation of Savonius-style rotors and have presented their work
in the published literature.3,4,35–39 The results which will be presented later in this report will
largely be drawn from this collection of work.
Heretofore, the discussion has focused on drag-based rotors that are fabricated by bisecting
cylinders. There have also been investigations of other geometries such as cup designs which
operate much like a wind-speed anemometer.38,39 Because of the predominance of cylindrical
rotor designs, they will be the exclusive focus of the remainder of this paper.
A detailed and comprehensive coverage of all major types of wind-power devices and anal-
ysis techniques is provided in Ref. 40.

EXPERIMENTS ON SAVONIUS TURBINES


Experiments on Savonius rotors take place either in field or with the use of a wind tunnel.
The vast majority of research described here has been completed using wind tunnels. It is a rar-
ity that detailed research occurs after installation.
With respect to wind tunnels, experiments are typically carried out by placing a scaled
rotor in the tunnel or by placing the rotor at the exit of a tunnel. Both methods have distinct
advantages and disadvantages. Among the critical issues is the knowledge of the wind speed
that passes the rotor. When a turbine is placed in a wind tunnel, there is a well-known blockage
effect that must be accounted for. The blockage effect is related to the fact that because of con-
servation of mass, the air must accelerate to move past the wind turbine. Consequently, the per-
centage of the wind tunnel cross section that is “blocked” by the turbine is manifested by an
acceleration of air. Since the available wind power varies with the cube of the wind speed,
even a small blockage ratio can lead to large impacts on the power coefficient.
042703-10 Abraham et al. J. Renewable Sustainable Energy 4, 042703 (2012)

For an object such as a Savonius turbine, the issue of blockage is complicated by the fact
that the blocked area changes as the device rotates so that no single value can be used. Further-
more, it is expected that rotational speed may impact the blockage effect, with a limiting value
for both a very rapidly rotating turbine and a very slowly rotating device. While a thorough
investigation of this dependency is currently being planned in wind tunnel tests, here an esti-
mate of the impact will be given with a justification of making the estimate.
Blockage can be dealt with by correcting for the oncoming wind speed or by derating the
power generated by the turbine. Both methods, should ultimately, provide similar results. Here,
the definition of the blockage correction factor F is the ratio of the power that would be pro-
duced by a wind turbine in a tunnel compared to a similar turbine in free space. Mathemati-
cally, this is expressed as

 3
Powertunnel Vtunnel
F¼  : (4)
Powerfree space Vfree space

The Savonius turbine tests by the present authors were completed in a large-scale wind tunnel.
The tests utilized a full-sized, single section of the Savonius rotor. It was decided to avoid
using a scaled turbine because the system was tested with the entirety of the accompanying
electronics. Coupling a scaled rotor with the electronic system was deemed inappropriate.
Wind speed measurements are made by calibrated instrumentation. Account was made of
the tunnel cross sectional area so that the wind speed at the turbine location was calculated.
The experiments required collection of varied data including electrical power, turbine rota-
tion (measured by optical tachometry), and wind speed. Tests were completed for a number of
wind speeds and for various rotor designs. The rotors were tested both with and without circular
plates which capped the edges of the blades. Additionally, the rotors were tested with and with-
out venting slots that were designed to reduce thrust loading on the tower while minimizing the
impact on power production. Figure 7 shows a photograph of the capped and vented rotor
which was tested in the wind tunnel.

FIG. 7. Photograph of the capped and vented rotor. Reproduced with permission from B. D. Plourde, J. P. Abraham, G. S.
Mowry, and W. J. Minkowycz, Wind Eng. 35, 213-220 (2011). Copyright 2011 Multi-Science Publishing Company.
042703-11 Abraham et al. J. Renewable Sustainable Energy 4, 042703 (2012)

FIG. 8. Cross section with dimensional nomenclature. Reproduced with permission from B. D. Plourde, J. P. Abraham, G.
S. Mowry, and W. J. Minkowycz, Wind Eng. 35, 213-220 (2011). Copyright 2011 Multi-Science Publishing Company.

For a given wind velocity, the power extracted from the wind by the combination of a
wind turbine and generator-with-load will have a maximum value. Two methods were used dur-
ing the wind tunnel tests to determine the power curve vs. load parameterized by wind speed.
In the first method, at each wind speed, the generator was connected to a resistive load that
could be easily varied. For our testing, load resistances ranging from 20 X to approximately
1000 X were used to determine the resulting power curve. In the second method, adaptive
power electronics with feedback was used to dynamically change the apparent load impedance
presented to the generator for maximum power extraction. Both methods lead to the same im-
pedance for any particular wind condition for extracting maximum power.
The cross section of the rotor is best displayed in Fig. 8 which shows an illustration of the
blade with dimensional nomenclature. It can be seen that this design differs from the more tra-
ditional cylindrical designs of prior experiments that are shown in Figs. 1 and 2. Among the
differences is the straight section that connects the two curved ends. Also, the curved ends are
not half-circles, as are required when the rotors are fabricated from bisecting a cylindrical
drum.
The relevant dimensions for Fig. 8 are shown in Table I.
Since the wind turbine occupied 23% of the wind tunnel cross section, the air speed had to
increase by approximately 23% at the test site. This increase in air speed is not identical to
23% for a number of reasons. First, the wind turbine is rotating so that on the advancing side,
the air that intersects the turbine frontal area is advancing, albeit at a lower speed. At the same
time, the air on the retreating side is actually moving upstream. Additionally, there is some
slight air passage through the slots, and it is also recognized that the projected area of the wind
turbine changes during its rotation. As a consequence of these facts, the use of the projected
area as a correction factor is made, with these uncertainties.
Using the relationship set forth in Eq. (4), this wind acceleration led to an 86% increase in
performance. Also, a review of the literature7,41 allowed an estimate of an increase in

TABLE I. Relevant dimensions for the turbine blade.

L1 0.92 m (36 in.)


L2 0.41 m (16 in.)
L3 0.10 m (4 in.)
R 0.20 m (8 in.)
h 120 deg
Height 1.1 m (43 in.)
042703-12 Abraham et al. J. Renewable Sustainable Energy 4, 042703 (2012)

FIG. 9. Power production from wind tunnel tests of a single-stage unvented, uncapped rotor. Reproduced with permission
from B. D. Plourde, J. P. Abraham, G. S. Mowry, and W. J. Minkowycz, Wind Eng. 35, 213-220 (2011). Copyright 2011
Multi-Science Publishing Company.

performance of approximately 100% for this blockage ratio. Consequently, power results were
de-rated by a factor of F (F ¼ 2). With this correction factor, the data shown in Figs. 9 and 10
result. Both figures present electrical power generation for a single-stage rotor. In Fig. 9, the
data are for an unvented rotor with no capping plate. Figure 10, however, includes both vents
and a circular cap. Additionally, the experiments were carried out for a number of electrical
loads which were used to simulate the electronics that would be used in practice.
A few conclusions can be drawn from the results. First, for the unvented and uncapped
case, there is a very weak dependence on electrical load. However, for the capped and vented
case, the performance is strongly linked to the electrical system. This finding suggests that the
electrical system should be designed in connection with the rotor. It also suggests that further
changes in electronics could have increased performance even further.

FIG. 10. Power production from wind tunnel tests of a single-stage vented and capped rotor. Reproduced with permission
from B. D. Plourde, J. P. Abraham, G. S. Mowry, and W. J. Minkowycz, Wind Eng. 35, 213-220 (2011). Copyright 2011
Multi-Science Publishing Company.
042703-13 Abraham et al. J. Renewable Sustainable Energy 4, 042703 (2012)

It was also found that in-field tests provided results that closely matched with those from
the wind tunnel.36 This close agreement, both in power production and in the rate of the rota-
tion of the turbine, strongly supports the quality of the results presented in Figs. 9 and 10.
In an effort to further increase performance, a new design was developed which had a
slightly larger vent compared to its counterpart of Fig. 8. It also had a slightly larger size (the
diameter was approximately 4 in. larger). The most significant difference is the redesign of the
circular cap which became contoured. The illustration in Fig. 11 shows a top-view of the rotor.
Presently, the authors are preparing to test the Fig. 11 device and results are forthcoming.
However, simulations have been performed on the newly designed rotor and results will be
shown later in this work.
The discussion of the experimental efforts here has focused on results of commonly
encountered Savonius rotors. A summary of the results is that a power coefficient that ranged
between 5% and 13% was achieved. These results are slightly lower than prior efforts but are
comparable. It is significant that the results presented here include losses in the electrical sys-
tem. When those losses are accounted for, the results set forth here are in closer agreement
with other researchers.

NUMERICAL SIMULATIONS
While traditionally, experimental techniques have formed the backbone of wind turbine
research, recent advances in simulation techniques have allowed a new vehicle for turbine
study.
Broadly, simulation studies are referred to as computational fluid dynamics or CFD which
allows a reconstruction of all properties of the fluid and its flow (pressure, density, velocity,
etc.) at all locations.
It is interesting to assess the advance of simulation methods that have occurred over the
very recent past. The first generation studies were strictly two-dimensional so that the simula-
tions were performed for a cross section of the rotor geometry.3,31 An additional restriction is
that some solutions were steady state.31,38 Importantly, the impact of the rotation on the fluid
motion itself was not fully accounted for.31,38
The justification of these two simplifications (two dimensional and steady state) is easily
seen in the context of computational power. A two-dimensional simulation requires far few ele-
ments (an order of magnitude or more) than does a three-dimensional calculation. In effect,
two-dimensional simulations treat the rotors as infinitely tall so that end-effects are unaccounted
for. The steady-state limitation means that a single solution of the fluid flow is needed to

FIG. 11. Top-view of a redesigned Savonius rotor with a contoured cap.


042703-14 Abraham et al. J. Renewable Sustainable Energy 4, 042703 (2012)

completely characterize the rotor performance for a given angular position. For a truly unsteady
solution, the fluid flow equations must be solved at many instances in time as the rotor revolves
about its axis.
To the best knowledge of the authors, only recently has a fully three-dimensional unsteady
simulation been carried out on a full-scale Savonius style turbine.35 To provide some context,
results of both two- and three-dimensional simulations will be given here and a comparison
with experimental results will also be shown.
Despite the novelty of the presented method, it is important to recognize the literature
which has been published on CFD for Savonius turbines. Among the relevant citations
unearthed by the authors are groups that deal with complex, sometimes three-dimensional flow
but either in a static environment or without proper accounting of rotational effects. Among the
references in this category are Refs. 42–45.
There is a larger body of literature that presents simulations on two-dimensional Savonius
rotors in a rotating environment, for instance, see Refs. 46–50. However, to the best knowledge
of the authors, the results presented in this article are the first from a fully three dimensional
unsteady calculation. In Ref. 51, a similar simulation was performed for an unsteady rotating
VAWT, however, it was of the Darrieus style.
The mathematics of the CFD method will not be discussed in great detail here but readers
are referred to Refs. 3 and 34 for more information.

TWO-DIMENSIONAL SIMULATIONS
The first step in the simulation effort is the generation of the computational mesh. The
two-dimensional mesh that was employed by the authors is shown in a series of increasing
magnifications in Fig. 12. A few characteristics are immediately apparent after viewing the
image. First, the elements encompassing the fluid are smaller in the vicinity of the rotor. Sec-
ond, there are very thin elements which are aligned with the surface of the rotor blade. These
elements are often called boundary-layer elements and are specially positioned to capture the
very steep gradients of velocity in the fluid at the surface of the rotor. Boundary-layer elements

FIG. 12. Increasing magnified views showing the computational mesh which subdivides the fluid region. Reprinted with
permission from J. P. Abraham, B. D. Plourde, G. S. Mowry, E. M. Sparrow, and W. J. Minkowycz, J. Renewable Sustain-
able Energy 3, 033109 (2011). Copyright 2011 American Institute of Physics.
042703-15 Abraham et al. J. Renewable Sustainable Energy 4, 042703 (2012)

are necessary for calculations such as this, particularly because the flow will be highly
turbulent.
The two-dimensional simulations of Ref. 3 utilized approximately 2 000 000 elements and
solved for four complete rotations of the turbine so that a periodic solution was guaranteed.
Each rotation was subdivided into 1000 time steps so 4000 solutions, in total, were generated.
The physical timestep was the time required for one rotation/1000. Since multiple rates of rota-
tion were investigated, the timesteps varied accordingly. The solutions were accomplished with
the ANSYS-CFX version 12.0 software.
The Reynolds numbers were greater than about 700 000 for the simulations and approach-
ing freestream turbulence levels were taken to be 5%. The shear stress transport turbulence
model was used,52 and a mesh refinement study was performed wherein the number of elements
was systematically increased until the results were mesh independent. The boundary-layer ele-
ments which are shown in Fig. 12 allowed yþ values of approximately 1 to be achieved.
Since the rate of revolution was not known a priori, it was necessary to assign a rotation
rate and to perform multiple calculations using different rates of rotation for each wind speed.
With this information available, it is possible to find the optimal performance of the turbine at
each wind velocity. In reality, when turbines are used in practice, it is possible to electrically
optimize them. The electrical load impacts the rate of rotation of the rotor so that a dynamically
varying electrical system can change as wind speeds are altered. Such a dynamic system can
adjust itself to ensure that the turbine is operating near its optimal performance.
Results from a two-dimensional simulation are considered an upper bound on the actual
performance of the device. The basis for this assertion is that in a two-dimensional calculation,
flow is not allowed to escape over or under the rotor blade, which would lead to power loss. A
comparison between the results from the two-dimensional simulations and the experiments al-
ready set forth is shown in Fig. 13. The dashed line is output from the simulations. Solid lines
are experimental results with different electrical loads.
An alternative way to compare simulated and experimental results is to show how power
generation varies with the tip-speed ratio. Such a comparison is shown in Fig. 14 which
presents information for a single wind speed. The comparisons correspond to the turbine shown
in Figs. 7 and 8 with dimensional details listed in Table I. It is seen that the simulations over-
predict the measured performance for all tip-speed ratios.

FIG. 13. Comparison of simulated and experimental results. Reprinted with permission from J. P. Abraham, B. D. Plourde,
G. S. Mowry, E. M. Sparrow, and W. J. Minkowycz, J. Renewable Sustainable Energy 3, 033109 (2011). Copyright 2011
American Institute of Physics.
042703-16 Abraham et al. J. Renewable Sustainable Energy 4, 042703 (2012)

FIG. 14. Comparison of simulated and experimental results for an approaching wind speed of 8.7 m/s. Results are for a
three-stage rotor system. Reprinted with permission from J. P. Abraham, B. D. Plourde, G. S. Mowry, E. M. Sparrow, and
W. J. Minkowycz, J. Renewable Sustainable Energy 3, 033109 (2011). Copyright 2011 American Institute of Physics.

One advantage of simulations over experimentation is that they can provide information
about the flow that is otherwise unavailable. For instance, local values of pressure and velocity
are available at all locations. Figure 15 has been prepared to show velocity-field information at
four instances in time for the two-dimensional calculations. In the images, flow is from the bot-
tom toward the top. The contours are scaled by the color legend on the left. It is seen that
downstream of the rotor, there is a large wake region. Upstream of the rotor the flow is unidir-
ectional and has a uniform velocity.
One significant outcome was that calculations performed with and without bounding walls
that simulate the walls of a wind tunnel showed that the positioning of the rotor in a wind

FIG. 15. Images of the velocity pattern in the fluid flowing around a rotating Savonius turbine. Reproduced with permission
from B. D. Plourde, J. P. Abraham, G. S. Mowry, and W. J. Minkowycz, Sens. Transducers J. 13, 53–61 (2011). Copyright
2011 International Frequency Sensor Association.
042703-17 Abraham et al. J. Renewable Sustainable Energy 4, 042703 (2012)

tunnel would increase performance by 95% compared to a turbine positioned in free space.
This finding, (F ¼ 1.95) is in close agreement with the two prior, independent methods of deter-
mining the blockage factor and adds further credibility to the calculations.

THREE-DIMENSIONAL SIMULATIONS
As mentioned earlier, the computational resources required to complete unsteady,
three-dimensional simulations are great. Nevertheless, such simulations are required in order to
properly account for flow that passes either above or below the rotor blades. Consequently, a
three-dimensional model of the contoured turbine was constructed and is shown in Fig. 11. A
schematic of the solution domain is shown in Fig. 16 with annotations calling out significant
features. It should be recognized that the turbine rotation is computed by placing it in a rotating
frame of reference that is surrounded by a stationary fluid frame. The image of Fig. 16 is taken
from a top-down view so that the three-dimensional nature of the problem is not readily
apparent.
The calculations encompassed four revolutions of the rotor; each revolution was subdivided
into 1000 time steps so that the physical timestep was the period of rotation/1000. It was found
that after the second revolution, the results were periodic. In total, 23 000 000 elements were
used in the simulation. As with the two-dimensional case, special boundary-layer elements were
positioned at all surfaces of the rotor so that values of yþ  1 were achieved. The diameter of
the rotor was 1.32 m and its height was 1.3 m. The calculations used 12 parallel processers with
3 GHz speed and 60 Gb of RAM. A single simulation for each wind speed required 45 days of
computer time.
An output of the calculation was the revolution-averaged torque, which, when multiplied
by the rate of rotation, provided the mechanical power. As was discussed with the
two-dimensional calculations, the rate of rotation was assigned a priori. In practice, the rate of
rotation is affected by wind speed and the electrical load. Because of this, it is possible to
dynamically alter the electrical system to ensure that the turbine continually functions near its
peak. A similar type of optimization can be accomplished numerically if solutions are carried
out for multiple rates of rotation and at multiple wind speeds as was discussed in the Two-
Dimensional Simulations section. This optimization process requires many months of calcula-
tions and is presently on-going. While the results are not yet available, the method that is
described here is straightforward. In the following, a discussion of the results of the first of the
three-dimensional unsteady simulations will be provided.

FIG. 16. Illustration of the top-view showing a contoured turbine positioned in a fluid environment. Reproduced with per-
mission from B. D. Plourde, J. P. Abraham, G. S. Mowry, and W. J. Minkowycz, Wind Eng. 35, 213–220 (2011). Copyright
2011 Multi-Science Publishing Company.
042703-18 Abraham et al. J. Renewable Sustainable Energy 4, 042703 (2012)

When a comparison was made between the power output from the simulated device with
the output obtained from experiments on similar sized, capped rotors, it was found that the
agreement was to within 10%. This level of agreement is excellent considering the complexity
of the problem and the difficulty of the calculations. It suggests that three-dimensional unsteady
simulations can now be considered a viable tool for design and optimization of Savonius-style
wind turbines. As computational speeds increase, the time needed to complete the simulations
required to fully characterize a Savonius turbine will decrease. Full details of the three-
dimensional simulations are provided in Ref. 35 but are omitted here for brevity.
Four advantages of numerical simulation are that: designs can be tested without the con-
struction of prototypes; full-scale models can be evaluated (wind tunnels usually require scaled
rotors); design alterations can be achieved relatively quickly to assess their impact on perform-
ance; and information about pressure and flow patterns can be obtained that otherwise would
not be possible through experimentation.
To provide some perspective on flow information from three-dimensional simulations, Fig.
17 has been prepared which shows flow patterns that are taken from a cutaway view through
the rotor cross section. The results correspond to an approaching wind speed of 8 m/s and a rate
of rotation of 60 rpm. In the image, the rotor is rotating in a counterclockwise direction. Some
key features can be observed. First, the flow accelerates as it moves around the rotor. This is
clearly evident by the shades of contour colors which are linked to the color legend to the left
of the figure. Second, there are separation zones on the backside of the rotor blades. There is
also fluid shown passing through the venting slots, particularly on the uppermost blade (return-
ing blade). The slots are designed to promote flow through the blade on the return side but not
on the advancing side. In this regard, it is clearly evident from the velocity patterns that the
vents are performing their intended duty.
The selection of the plane to view flow patterns was made almost arbitrarily in Fig. 17; the
figure is intended to be representative. Other images of velocity patterns at other locations
could also have been created. Also, the angular position that was selected for Fig. 18 was arbi-
trary. Flow patterns can be found at any position during the rotating motion. The one selected
here is meant to be illustrative.
Another way to view flow patterns is through streamline plots. Figure 18 has been prepared
to show streamline patterns on two different planes. The first plane, shown in (a), is near the
center of the rotor whereas the second plane (b) is at the top of the rotor. The flow is moving
left-to-right and the rotor is spinning counterclockwise. It is seen that in the middle of the rotor,
the fluid vectors show a wide divergence around the spinning blades. On the other hand, near
the top of the rotor, the streamlines are seen to pass over the contoured cap which is shown in
grey. This type of flow bypass is not able to be captured in two-dimensional simulations.

FIG. 17. Velocity patterns for flow through a slotted, three-dimensional turbine. Results are shown on a plane which passes
through the center of the rotor.
042703-19 Abraham et al. J. Renewable Sustainable Energy 4, 042703 (2012)

FIG. 18. Streamline patterns extracted at two horizontal planes.

CONCLUDING REMARKS
This paper has attempted to serve multiple purposes. First, we have compared the various
styles of wind turbines and discussed the advantages each style has which may make it suitable
for particular applications. Motivated by the specific applications that befit Savonius-style tur-
bines, we have provided a review of the relevant literature. That review, though not exhaustive,
does capture most of the critical contributions to the field and it allows the chronology of Savo-
nius turbine research to be seen.
Second, a summary of various experimental results of the performance of Savonius turbines
was given which showed that while there is variance in the studies, the values of the power
coefficient which is most commonly reported is in the 0.10-0.15 range. In some cases, the effi-
ciencies are significantly larger than this, however, the prevailing consensus is that the peak
performances is approximately 0.15. Differences in the literature may be attributed to experi-
mental technique, including whether blockage effects are included or whether frictional or elec-
trical losses are accounted for. It is also possible that the dimensional value of the rotor radius
has an impact on the power coefficient which is not apparent when results are presented in
dimensionless form.
Third, a discussion of recent advances in numerical simulation was provided which has
shown that the advancement of simulation methods has allowed Savonius calculations to evolve
from steady, two-dimensional, laminar flow simulations to unsteady, fully three-dimensional
calculations with full account of turbulent fluid motion. Presently, the time required to complete
042703-20 Abraham et al. J. Renewable Sustainable Energy 4, 042703 (2012)

each calculation is significant, however, with projected increases in computational performance


and parallel processing, it is likely that computational studies will be able to be completed far
quicker than prototype studies in the very near future. Results from both two- and three-
dimensional simulations were provided. Two-dimensional calculations assume, in essence, that
the turbines have an infinite height. Consequently, results from two-dimensional simulations are
an upper bound on the actual performance of a real turbine.
The simulations and experiments presented here were in good agreement. Furthermore, the
data presented here showed the positive impact of capping which allows significant improve-
ments in the power coefficient. Vents were also employed to lower thrust loading on the tower.

ACKNOWLEDGMENTS
The authors gratefully acknowledge financial support from Windstrip, LLC.
1
See http://www.gwec.net/fileadmin/images/News/Press/GWEC_-_Global_Wind_Statistics_2011.pdf for Global Wind
Statistics 2011 by Global Wind Energy Council.
2
J. O. Dabiri, “Potential order-of-magnitude enhancement of wind farm power density via counter rotating vertical-axis
wind turbine arrays,” J. Renewable Sustainable Energy 3, 043104 (2011).
3
J. P. Abraham, B. D. Plourde, G. S. Mowry, E. M. Sparrow, and W. J. Minkowycz, “Numerical simulation of fluid flow
around a vertical-axis turbine,” J. Renewable Sustainable Energy 3, 033109 (2011).
4
U.S. Department of Energy, “20% Wind Energy by 2030,” in Meeting the Challenges, Arlington, VA, 6-7 October 2008.
5
B. D. Plourde, J. P. Abraham, G. S. Mowry, and W. J. Minkowycz, “Use of small-scale wind energy to power cellular
communication equipment,” Sens. Transducers J. 13, 53–61 (2011).
6
S. J. Savonius, “The S-rotor and its applications,” Mech. Eng. 53, 333–337 (1931).
7
A. J. Alexander, “Wind tunnel corrections for savonius rotors,” in 2nd International Symposium on Wind Energy Sys-
tems, London, 3-6 October 1978.
8
A. J. Alexander and B. P. Holownia, “Wind tunnels tests on a savonius rotor,” J. Wind Eng. Ind. Aerodyn. 3, 343–351
(1978).
9
R. E. Sheldahl, B. F. Blackwell, and L. V. Feltz, “Wind tunnel performance data for two- and three-bucket savonius
rotors,” J. Energy 2, 160–164 (1978).
10
A. T. Sayers, “Blade configuration optimization and performance characteristics of a simple savonius rotor,” Proc. Inst.
Mech. Eng., Part A 199, 185–191 (1985).
11
I. Ushiyama and H. Nagai, “Optimum design configurations and performance of Savonius rotors,” Wind Eng. 12, 59–75
(1988).
12
V. J. Modi and M. S. U. K. Fernando, “On the performance of the Savonius wind turbine,” J. Sol. Energy Eng. 111,
71–81 (1989).
13
T. Ogawa and H. Yoshida, “Development of rotational speed control systems for a Savonius-type wind turbine,” J. Fluids
Eng. 111, 53–58 (1989).
14
N. Fujisawa and F. Gotoh, “Pressure measurements and flow visualization study of a Savonius rotor,” J. Wind Eng. Ind.
Aerodyn. 39, 51–60 (1992).
15
N. Fujisama and F. Gotoh, “Experimental study on the aerodynamic performance of a Savonius rotor,” J. Sol. Energy
Eng. 116, 116–149 (1994).
16
A. S. Grinspan, “Development of a low speed wind tunnel and testing of Savonius wind turbine rotors with twisted
blades,” Master of Technology (Indian Institute of Technology, Guwahati, 2002).
17
J. R. K. Mulakalapalli, “Experimental investigation and flow simulation of Savonius rotors,” Master of Technology (In-
dian Institute of Technology, Guwahati, 2004).
18
F. White, Viscous Fluid Flow, 3rd ed. (McGraw-Hill, New York, NY, 2006).
19
E. M. Sparrow, J. C. K. Tong, and J. P. Abraham, “Fluid flow in a system with separate laminar and turbulent zones,”
Numer. Heat Transfer, Part A 53, 341–353 (2008).
20
J. P. Abraham and A. P. Thomas, “Induced co-flow and laminar-to-turbulent transition with synthetic jets,” Comput. Flu-
ids 38, 1011–1017 (2009).
21
J. P. Abraham, J. C. K. Tong, and E. M. Sparrow, “Breakdown of laminar pipe flow into transitional intermittency and
subsequent attainment of fully developed intermittent or turbulent flow,” Numer. Heat Transfer, Part B 54, 103–115
(2008).
22
J. P. Abraham, E. M. Sparrow, and J. C. K. Tong, “Heat transfer in all pipe flow regimes—Laminar, transitional/intermit-
tent, and turbulent,” Int. J. Heat Mass Transfer 52, 557–563 (2009).
23
W. J. Minkowycz, J. P. Abraham, and E. M. Sparrow, “Numerical simulation of laminar breakdown and subsequent inter-
mittent and turbulent flow in parallel plate channels: Effects of inlet velocity profile and turbulence intensity,” Int. J. Heat
Mass Transfer 52, 4040–4046 (2009).
24
R. D. Lovik, J. P. Abraham, W. J. Minkowycz, and E. M. Sparrow, “Laminarization and turbulentization in a pulsatile
pipe flow,” Numer. Heat Transfer, Part A 56, 861–879 (2009).
25
J. P. Abraham, E. M. Sparrow, J. C. K. Tong, and D. W. Bettenhausen, “Internal flows which transist from turbulent
through intermittent to laminar,” Int. J. Therm. Sci. 49, 256–263 (2010).
26
A. P. Thomas and J. P. Abraham, “Sawtooth vortex generators for underwater propulsion,” Open Mech. Eng. 4, 1–7
(2010).
27
J. P. Abraham, E. M. Sparrow, and W. J. Minkowycz, “Internal-flow Nusselt numbers for the low-Reynolds-number end
of the laminar-to-turbulent transition regime,” Int. J. Heat Mass Transfer 54, 584–588 (2011).
042703-21 Abraham et al. J. Renewable Sustainable Energy 4, 042703 (2012)

28
T. Gebreegziabher, E. M. Sparrow, J. P. Abraham, E. Ayorinde, and T. Singh, “High-frequency pulsatile pipe flows
encompassing all flow regimes,” Numer. Heat Transfer, Part A 60, 811–826 (2011).
29
J. P. Abraham, E. M. Sparrow, W. J. Minkowycz, R. Ramazani-Rend, and J. C. K. Tong, All Fluid-Flow-Regimes Simula-
tion Model for Internal Flows (Nova Science, New York, 2010).
30
J. P. Abraham, E. M. Sparrow, W. J. Minkowycz, R. Ramazani-Rend, and J. C. K. Tong, “Modeling internal flows by an
extended menter transition model,” in Turbulence: Theory, Types, and Simulation (Nova, New York, 2011).
31
M. J. Rajkumar, U. K. Saha, and D. Maity, “Simulation of flow around and behind a Savonius rotor,” Int. Energy J. 6,
83–90 (2005).
32
M. J. Rajkumar and U. K. Saha, “Valve-aided twisted Savonius rotor,” Wind Eng. 30, 243–254 (2006).
33
U. K. Saha and M. J. Rajkumar, “On the performance analysis of Savonius rotor with twisted blades,” Renewable Energy
31, 1776–1788 (2006).
34
U. K. Saha, S. Thotla, and D. Maity, “Optimum design configuration of Savonius rotor through wind tunnel experiments,”
J. Wind Eng. Ind. Appl. 96, 1359–1375 (2008).
35
B. D. Plourde, J. P. Abraham, and G. S. Mowry, “Simulation of three-dimensional vertical-axis turbines for communica-
tion applications,” Wind Eng. (in press).
36
B. D. Plourde, J. P. Abraham, G. S. Mowry, and W. J. Minkowycz, “An experimental investigation of a large, vertical-
axis wind turbine: Effects of venting and capping,” Wind Eng. 35, 213–220 (2011).
37
J. P. Abraham, B. D. Plourde, and G. S. Mowry, “Fluid dynamic simulations of wind turbines,” in ANSYS Regional Con-
ference, Minneapolis, MN, 20 October 2011.
38
G. S. Mowry, R. A. Erickson, and J. P. Abraham, “Computational model of a novel two-cup horizontal wind-turbine sys-
tem,” Open Mech. Eng. J. 3, 26–34 (2009).
39
J. P. Abraham, G. S. Mowry, and R. A. Erickson, “Design and analysis of a small-scale vertical-axis wind turbine for
rooftop power generation,” in Climate Change Technology Conference, Hamilton, Ontario, 12-15 May 2009.
40
R. E. Wilson and P. B. S. Lissaman, “Applied aerodynamics of wind power machines,” NSF Research Report, May 1974;
available online at http://ir.library.oregonstate.edu/xmlui/bitstream/handle/1957/8140/WilsonLissaman_AppAeroOfWindPwr
Mach_1974.pdf;jsessionid=F5D355A47A1DC63F5D699D3174D7C398?sequence=4.
41
A. Biswas, K. Gupta, and K. K. Sharma, “Experimental investigation of overlap and blockage effects on three bucket
Savonius rotors,” Wind Eng. 31, 363–368 (2007).
42
B. Altan and M. Atilgan, “An experimental and numerical study on the improvement of the performance of Savonius
wind rotor,” Energy Convers. Manage. 49, 3425–2432 (2008).
43
O. B. Yaakob, K. B. Tawi, and D. T. S Sunanto, “Computer simulation studies on the effect of overlap ratio for Savonius
type vertical axis marine current turbine,” Int. J. Eng. Trans. 23, 79–88 (2010).
44
T. Li, J. R. Chai, F. N. Dang, and W. Q. Tian, “Numerical simulation on static performance of Savonius rotor based on
different turbulence models,” Adv. Mater. Res. 488–489, 1238–1242 (2012).
45
P. Jaohindy, “Aerodynamic and mechanical system modeling of a vertical axis wind turbine (VAWT),” in 2011 Interna-
tional Conference on Electrical and Control engineering, Le Tampon, France, 16-18 September 2011.
46
I. Dobrev and F. Massouh, “CFD and PIV investigation of unsteady flow through Savonius wind turbine,” Energy Proce-
dia 6, 711–720 (2011).
47
T. Zhou and D. Dempfer, “Numerical study for detailed flow fields and performance of the Savonius-type rotor,” in Bul-
letin of the American Physical Society Annual Meeting, Baltimore, MA, 20-22 November 2011.
48
R. Ricci, S. Montelpare, A. Secchiaroli, and V. D’Alessandro, “Flow field assessment in a vertical axis wind turbine,”
Adv. Fluid Mech. VIII, WIT Trans. Eng. Sci. 69, 255–267 (2010).
49
M. H. Mohammed, G. Janiga, E. Pap, and D. Thecenin, “Optimization of Savonius turbines using an obstacle shielding
of the return blade,” Renewable Energy 35, 2618–2626 (2010).
50
V. D’Alessandro, S. Montelpare, R. Ricci, and A. Secciaroli, “Unsteady aerodynamics of a Savonius wind rotor: A new
computational approach for the simulation of energy performance,” Energy 35, 3349–3363 (2010).
51
R. Howell, N. Qin, J. Edwards, and N. Durrani, “Wind tunnel and numerical study of a small vertical axis wind turbine,”
Renewable Energy 35, 412–422 (2010).
52
F. Menter, “Two-equation Eddy-viscosity turbulence models for engineering applications,” AIAA J. 32, 1598–1605
(1994).

View publication stats

You might also like