You are on page 1of 42

Chapter 4.

Hydrokinetic turbines blade design

Chapter 4

Hydrokinetic turbines blade


design

4.1. Introduction
This chapter describes the design of hydrokinetic turbine blades to be used in a
horizontal axis hydrokinetic turbine prototype generating free and clean energy for
remote communities in Sarawak, Malaysia. Relevant recommendations resulting from
case studies studied in Chapter 3 will be implemented in this chapter. However, before
covering the actual blade design procedures, the approach and method used in blade
design for this research are first introduced. Their validity in hydrokinetic turbine blade
design is explained. The performance characteristics for various turbines are revisited
and the choice of turbine for this research is further justified. This chapter then proceeds
to describe basic principles and the use of relevance formulas/equations in the design of
a hydrokinetic turbine blade. Optimum blade design procedures, blade coordinates
calculations and twist angles calculations are explained and demonstrated. A wooden
blade template making is also explained and the used of this template in making three
identical blades for a turbine rotor is demonstrated. Finally issues regarding wood
blades are explained briefly.

4.2. Validity of wind turbines design principles for


hydrokinetic turbines
The approach used in the design of a hydrokinetic turbine prototype in this research is
similar to that of wind turbines. Adopting wind turbines design principles is commonly
an accepted approach in hydrokinetic turbine design, mainly because water and air are
both fluids, but with significant different in density [4.1], [4.2] and [4.3]. Therefore the
power equation (2.1) stated in Chapter 2 section 2.4.1 and blade element momentum
theory (BEM theory) developed for wind turbines design are generally applicable for
hydrokinetic turbine for as long as cavitation is avoided [4.4], [4.5]. In addition, the

Water current energy for remote community: Design and testing of a clog-free horizontal axis 85
hydrokinetic turbine
Chapter 4. Hydrokinetic turbines blade design

application of wind turbines’ BEM theory in hydrokinetic turbine design has been well
validated by a number of publications. For example, wind turbine performance code
WT-Perf based on BEM theory was reported to have good accuracy in modeling a 35
kW Verdant power Gen4 turbine installed in East River, New York [4.6]. The GH-Tidal
and SERG-Tidal codes, which are wind turbines’ BEM theory based computer
programs, enhanced for hydrokinetic turbines design were also reported having
satisfactory prediction accuracy when applied to an 800 mm diameter laboratory scale
horizontal axis hydrokinetic turbine tested in a towing tank [4.7]. Therefore, based on
the argument stated above, the design of a hydrokinetic turbine prototype intended in
this research will be confidently based on wind turbines’ BEM theory, keeping in mind
that cavitation must be avoided.

4.3. Hydrokinetic turbine versus wind turbine


Despite the similarities with wind turbines, hydrokinetic turbines actually have some
significant differences, especially in terms of working environments, physical features
and challenges they are facing. The following subsections outline important
dissimilarities that must be noted in designing a hydrokinetic turbine.

4.3.1. Water density versus wind density


Perhaps one obvious dissimilarity is fluid density. The density of fresh water is 1000
kg/m3 (about 1025 kg/m3 for surface seawater) whereas air density is 1.223 kg/m3 [4.3].
This implies that water is more than 800 times denser than air. Because of this
significant difference, hydrokinetic turbine design reserves some considerations when
adopting wind turbine design approach. For example turbine operation in high density
water may give rise to cavitation which may disqualify the use of power equation and
BEM theory adopted from wind turbines design as mentioned in section 4.2.

The high water density also causes some forces experience by hydrokinetic turbines’
blades to differ. In wind turbines sector, the blades are usually long, thin, light and
slender in order to capture as much energy as possible from low density wind.
According to [4.8], as the blades spin fast, centrifugal loads dominate which tend to
restrict bending. This may help to prevent the blades from snapping. However, blades
used in hydrokinetic turbines are usually short, thick and made from selected materials
to ensure enough strength to resist bending loads.

Water current energy for remote community: Design and testing of a clog-free horizontal axis 86
hydrokinetic turbine
Chapter 4. Hydrokinetic turbines blade design

However, the high water density has an important advantage. Energy content in water
current is much higher even though usually having low velocity. This allows the design
of hydrokinetic turbines to be compact and should be easier to handle yet produce much
power as huge wind turbines. For a comparison, a 24 m diameter hydrokinetic turbine
operating in a 2 m/s water current produces power equal to a massive 60 m diameter
wind turbine operating in 7 m/s wind [4.4].

4.3.2. Strong water currents versus strong winds


As wind turbines are designed to withstand strong gusty winds such as hurricane,
tornados etc., hydrokinetic turbines must be designed to withstand strong water current
too. Attention must be observed in the design of hydrokinetic turbine operating in rural
rivers where flashfloods can occur easily, causing strong current and carry massive
amount of debris downriver. Two evidences of serious implication of intense force
created by river current are shown in Figure 4.1 and Figure 4.2 [4.9]. Figure 4.1 is
actually a snapshot taken from New York Times online video (URL cannot be retraced).
This figure shows two of the turbine’s blades were snapped at the middle probably due
strong bending force by strong current. On the other hand, Figure 4.2 shows turbine’s
blades have detached from their hub due to strong downriver current. According to
[4.9], the blades were replaced with strong aluminum-magnesium alloy, enabling the
turbine to withstand strong bending forces throughout the remaining test duration in
East River, New York.

Figure 4.1. Broken hydrokinetic turbine blades [New York Time Video 2009]

Water current energy for remote community: Design and testing of a clog-free horizontal axis 87
hydrokinetic turbine
Chapter 4. Hydrokinetic turbines blade design

Figure 4.2. Detached hydrokinetic turbine blades due to strong water current [4.9]

4.3.3. Clogging by debris


Both horizontal and vertical axis turbines are susceptible to clogging by river debris
such as leaves, tree branches, small trees, weeds etc. The explanation on the effect of
river debris and evidences of turbines clogged by debris are given in Chapter 2, sections
2.6.3.1 and 2.6.3.2. Besides impacting and clogging turbines’ blades, logs and
driftwoods were also found to affect turbines’ pontoon and anchor chains or cables. An
evidence of this is shown in Figure 4.5 [4.10], where hips of driftwoods have pileup on
the turbine’s pontoon and its anchor cable.

In fact the case studies presented in Chapter 2 suggests that debris problem is a major
limiting factor for river hydrokinetic turbines, preventing their widespread use in remote
communities’ application. This problem also could partly have shifted developers’
attention to huge marine current turbines such as the case reported in [4.11] where a
river turbines pioneer moved to develop massive marine current turbines.

Compared to wind turbines, clogging problem does not exist. Besides fouling by insects
and dust wind, turbines may occasionally hit birds in the air but this shouldn’t cause
abrupt and frequent power interruption as anticipated for stand-alone hydrokinetic
turbines supplying power to remote villages.

Water current energy for remote community: Design and testing of a clog-free horizontal axis 88
hydrokinetic turbine
Chapter 4. Hydrokinetic turbines blade design

Figure 4.3. Driftwoods piled up on hydrokinetic turbine pontoon and anchor cable [4.10]

4.3.4. Low water current speed


The operating water current speed for hydrokinetic turbines is known to be significantly
lower than the operating wind speed for wind turbines. According to [4.3],
hydrokinetic turbines may be designed for effective water current velocities ranging
from 1.75 to 2.25 m/s compared to 11 to 13 m/s for wind turbines. However, one
hydrokinetic turbine reported in Chapter 2 reference [2.16] can operate and able to
produce useful power in water current velocity as low as 0.5 m/s. Due to the low current
velocity, hydrokinetic turbines is expected to have much lower revolution per minute
(rpm) compared to that of wind turbines. For comparison, a two meter diameter
horizontal axis hydrokinetic turbine reported in [2.21] achieved 70 rpm in 2.1 m/s water
current, requiring 1:14 speed up gear ratio in order to drive a low speed (600 rpm)
permanent magnet generator but small wind turbines can operate at hundreds of rpm
[4.5], enabling some to direct drive (without speed up gears) low speed generators.

On the plus side, water currents are more constant and predictable compared to wind.
Unlike wind, water currents’ velocity and flow rate may take hours, days, weeks,
months or a season to change. This potentially assists in planning and operation matters,
and also allows hydrokinetic turbine to harvest more energy per year than wind
turbines.

Water current energy for remote community: Design and testing of a clog-free horizontal axis 89
hydrokinetic turbine
Chapter 4. Hydrokinetic turbines blade design

4.3.5. Stall and cavitation


Both wind and hydrokinetic turbines are designed to operate efficiently and safely in a
specific range of fluid velocities. Any operation above the design velocities will result
in a phenomenon called stall, happens in both hydrokinetic and wind turbines. Stall
decreases turbines’ output performance as fluids (water or wind) start to separate from
the surface of the turbines’ blades. While stall occurrence in wind turbines can be
utilized for good use i.e. to regulate output power of wind turbines [4.12], stall must be
avoided or delayed in hydrokinetic turbines in order to avoid cavitation [4.6]. This is
because, besides losing power due to hydrodynamic inefficiency (resulting from flow
separation), the impact of continuous built up and collapsing of air bubbles during
cavitation damages blades surface and further reduce turbines’ performance [4.5].

4.3.6. Corrosion, fouling, aeration


Other dissimilarities that hydrokinetic turbines have in comparison to wind turbines
include corrosion, bio fouling and aeration due to insufficient depth. Extreme moisture
and underwater environment are well understood to cause serious degree of corrosion
compared to on land environment where wind turbines operate. Turbines’ blades and
structures have to be specially treated or made from selected materials to avoid
corrosion [4.8].

In marine hydrokinetic turbines, fouling or marine growth on turbine blades and


structures can be a major problem. For example a detail study found in [4.13], indicates
that bio fouling can decrease turbines performance from 20% to 70%. This is serious
adversity compared to the effect of fouling by dust and insect carcasses in wind
turbines.

The problem of ventilation or aeration is unique to hydrokinetic turbines where air is


sucked by blades tips operating at high speed near to the water surface. This problem is
known to reduce turbines’ performances [4.16] and induce cavitation [4.2]. An example
of ventilation problem affecting performance of a vertical axis hydrokinetic turbine is
described in [4.14]. The ventilation problem could be a significant issue for turbines
operating in shallow rivers in remote areas where rivers are usually limited in depth and
width.

Water current energy for remote community: Design and testing of a clog-free horizontal axis 90
hydrokinetic turbine
Chapter 4. Hydrokinetic turbines blade design

Finally, Table 4.1 summarizes the arguments (presented above), highlighting


dissimilarities between hydrokinetic turbines and wind turbines.

No. Hydrokinetic turbines Wind turbines

3 3
1. Density of freshwater = 1000 kg/m . Density of air =1.223 kg/m .
Density of Seawater (surface) = 1025
3
kg/m .
Therefore water is more than 800
times denser than water.

2. Short , thick/fat and strong blades. Long and slender blades.


Bending loads dominate, tend to Centrifugal loads dominate, tend to restrict bending
snap or detached turbine blades. on blades.

3 Have to withstand strong current Have to withstand strong and gusty winds especially
during flashflood. during cyclone, tornado, hurricane etc.

4. Prone to clogging by river debris. No clogging.


Foul by waterborne plants and Foul by insects.
organisms.

5 Prone to impact by logs. Occasionally impact by birds and hail stones.

6. Operate at low water current speeds Operate at high wind speed up to 13 m/s.
of 0.5 - 2.25 m/s.

7. Stall reduces turbines’ power output Stall reduces turbines’ output power.
and causes cavitation. Stall can be used to regulate output power.
Cavitation must be avoided.

8. Aeration/ventilation causes reduction Not exist.


in power output.

9. Turbines and structures require Corrosion and fouling are not serious issues.
selected materials and must be
treated to avoid corrosion and bio
fouling.

10. Turbines’ sizes are limited by dept Not an issue.


and width of river

11. Operate at low rpm. Operate at higher rpm.


Require large step up gear ratio. Require lesser step up gear ratio.

Table 4.1. Dissimilarities between hydrokinetic turbines and wind turbines

Water current energy for remote community: Design and testing of a clog-free horizontal axis 91
hydrokinetic turbine
Chapter 4. Hydrokinetic turbines blade design

4.4. Adopting wind turbines design principles in


hydrokinetic turbines design.
The arguments presented in section 4.2 indicate that wind turbine design principles are
valid and can be used in hydrokinetic turbines design, provided that cavitation must be
avoided. Therefore, before advancing into the actual blade design, a brief review of
wind turbines’ characteristics is required in order to reaffirm hydrokinetic turbine of
choice for this research and properly select its basic features.

4.4.1. Cp comparison for various turbines


Figure 4.4, taken from [4.15] shows a comparison of rotor coefficients of power, Cp for
various wind turbines. Turbines presented in this figure, (even though wind turbines but
also apply to hydrokinetic turbines) can be categorized into two main groups, namely
the horizontal axis turbine and the vertical axis turbine. The terms horizontal and
vertical refer to the direction of fluids flow in reference to turbines’ axial.

Notice in this figure that the propeller type of turbine belong to the horizontal axis
category is slightly higher in efficiency compared to the Darrieus turbine which belong
to the vertical axis turbine. Vertical axis turbines used in rivers are usually straight
blades or H-Darrieus and not the original “egg-beater” shape Darrieus turbine because
the “egg-beater” shape has no advantage in water where fluid dynamic forces exceed
inertial forces.

Compared to the Savonius rotor which is a drag turbine (including Persian wheel and
waterwheel) , both the horizontal and vertical axis turbines which are lift type of
turbines operating on lift force, resulting from the effect of fluid flow around optimum
blades’ profile are significantly higher in efficiency.

Water current energy for remote community: Design and testing of a clog-free horizontal axis 92
hydrokinetic turbine
Chapter 4. Hydrokinetic turbines blade design

Figure 4.4. The power coefficient of various types of turbines [4.15]

By analyzing Figure 4.4, a three bladed propeller or horizontal axis turbine achieves the
highest Cp. According to [4.15], a modern propeller turbine used in wind power
systems can achieve power coefficient near 0.5 which is close to the maximum
theoretical limit or Betz limit (0.59) whereas other turbines have Cp around 0.4 and
below. For a small hydrokinetic turbine operating in very low water current speed, in it
is important to have the highest Cp in order produce a useful amount of power.

Two bladed horizontal axis turbines may be cheaper and easier to built, but the fact that
it requires higher fluid velocity to produce maximum output is a disadvantage for
hydrokinetic turbine operating in dominantly slow river current. On the other hand,
three bladed wind turbines are known to produce smoother output torque compared to
two blades turbine. It may be argued that hydrokinetic turbines output whether three or
two bladed may not experience serious fluctuation in water, but to be on the save side, a
three bladed lift type horizontal axis turbine is used in this research, in parallel with
recommendation made in Chapter 3. Difference in flow velocity at difference depth of a
river and turbine drive leg shadow similar to wind shear and tower shadow that cause
output fluctuation in wind turbines do exist in hydrokinetic turbine operating
environment. Smooth torque is required in electric generating turbines so that the output
can be directly and safely delivered to consumers.

Water current energy for remote community: Design and testing of a clog-free horizontal axis 93
hydrokinetic turbine
Chapter 4. Hydrokinetic turbines blade design

4.4.2. Power equation


Since wind turbine design method is applicable for hydrokinetic turbines, the governing
equation for turbine power calculation is the same as equation (2.1), repeated below for
ease of explanation.

The fluid density ⍴ value for water is much higher that air which is 1000 kg/m3
compared to 1.223 kg/m3 for wind (see table 4.1). The limit to how much power can be
harnessed from a moving fluid is 59% of the kinetic energy flux passing through the
blades in free flow. This happens (in theory) when the incoming fluid velocity is slowed
down to a third by an unshrouded turbine rotor. This limit is called Betz limit or
maximum theoretical coefficient of power Cp and it applies to both lift-type wind and
hydrokinetic turbines. However, the actual achievable Cp is lower than the Betz limit
because of losses due to hydrodynamic imperfections of blades and the efficiency of the
drive trains (generator, shaft and bearings etc.). For example, the maximum Cp reported
for a small horizontal axis hydrokinetic turbine is about 46 % [4.16]. Due to losses in
drive trains, the water-to-wire efficiency reported for hydrokinetic turbines is usually
less than 30% (see table 3.1 and 3.2 in Chapter 3).

Referring to the power equation (2.1), the cubic relationship between velocity and
power indicates that the output of a hydrokinetic turbine is highly dependent on water
current velocity. For example, if water current velocity is doubled, the output power
will increase eight times. However, in a real situation, one does not have control over
water current velocity. Therefore designers are left with the options of altering the size
of turbine, choosing a turbine with the highest Cp and using efficient power trains in
order to arrive at the required output.

4.4.3. Influence of blade number on Cp


The number of blades used on a turbine rotor and their chord length determines the
solidity, affects the fluid dynamic efficiency, Cp of its rotor and shifts the tip speed ratio
(tsr) at which maximum efficiency is achieved. For example, study in wind turbines
[4.15] shows that about 2% increase in power was achieved when blade number was
increased from 2 to 4 blades. While the increase in power may not be significant, the

Water current energy for remote community: Design and testing of a clog-free horizontal axis 94
hydrokinetic turbine
Chapter 4. Hydrokinetic turbines blade design

shift in tip speed ratio tsr where maximum Cp occurs is quite significant. For example,
the tsr in Figure 4.5 increases from about 6.5 to 10.5 when the blade number is
decreased from 4 to 2.

Turbines with higher tsr can be advantages for wind turbines (low weight, less cost and
high rpm) but achieving higher optimum tsr for hydrokinetic turbines could be an issue
because water currents velocity is usually very low. Not meeting the optimum tsr means
that the turbine is operating at below rated power, therefore not utilizing turbine’s full
capacity.

Another possible drawback of operating turbine at high tsr is cavitation inception. To


avoid this problem, the tip speed for hydrokinetic turbines is limited to less than 12 m/s
[4.11] or tip speed ratio below 7 [4.16]. However, small hydrokinetic turbines usually
attain optimum output at tsr about 3. See [4.17] and [4.18].

Figure 4.5. Influence of blade number on Cp and tsr [4.15]

As stated in section 4.4.1, using two bladed turbines may seem to be a good choice for a
community turbine due to cost constraint, but having to few blades may cause
significant pulsating or flickering electrical outputs, unsuitable for direct AC
transmission to consumers. Furthermore, uneven torque may cause a turbine to shake
and vibrate which may leads to stress in its structures and other components. Therefore

Water current energy for remote community: Design and testing of a clog-free horizontal axis 95
hydrokinetic turbine
Chapter 4. Hydrokinetic turbines blade design

a 3-bladed rotor is regarded as a fair compromise and was adopted, developed and
tested in this research.

4.4.4. Influence of rotor solidity on blade chord and tsr


Rotor solidity is defined as the ratio of total blade area over rotor swept area and it rotor
characteristics and its operating tsr. Figure 4.6 shows that the shape of optimum wind
turbine blades becomes more slender as more blades are used to make a rotor operating
at a given tsr. In reference to Figure 4.6, as the solidity decreases (less blade area over
swept area), the design tsr or λ increases. While high speed is a favorable feature of
electric generating turbines, the extreme slenderness of the blades may not have enough
strength and stiffness in hydrokinetic turbine application. This adds to the justification
of choosing a 3-bladed rotor and tsr equal to 4 for a turbine prototype in this research.

Figure 4.6. Blade chord decreases as number of blades and tsr increase [4.15]

A useful hind of a practical rotor solidity for small hydrokinetic turbine suitable for
remote communities can be found in [4.19]. According to this reference, more than one
horizontal axis hydrokinetic turbines with different degree of solidities were tested in
Brazil and a rotor solidity of about 30% was found to be superior. Another clue is

Water current energy for remote community: Design and testing of a clog-free horizontal axis 96
hydrokinetic turbine
Chapter 4. Hydrokinetic turbines blade design

found in [4.18] where a small horizontal axis hydrokinetic turbine having rotor solidity
of about 20% tested in a towing tank was found to archive high Cp of 45%.

4.4.5. Torque versus solidity


Figure 4.7 [4.15] shows that torque produced by wind turbines (also apply to
hydrokinetic turbines) is inversely related to its solidity. In reference to the figure given,
the American-style multi blade (high solidity) turbine produces very high torque
compared to a 2-bladed (low solidity) turbine therefore very successful in high torque
application such as water pumping in farms across America. However, turbines with
narrow and less blades can operate at much higher speed making them more suitable for
driving electric generators which requires high rpm.

The low starting torque problem in wind turbines is normally solved by running the
generator as a motor in order to turn the turbine to a desired speed before switching it
back to generator mode. However starting torque may not be a major problem in a
hydrokinetic turbine but over speed may cause cavitation and aeration problems. It is
also important to make sure that the blades used in a hydrokinetic turbine rotor are not
too narrow and slender in order ensure adequate strength and stiffness to withstand
strong water currents.

Figure 4.7. Influence of solidity of rotor torque [4.15]

Water current energy for remote community: Design and testing of a clog-free horizontal axis 97
hydrokinetic turbine
Chapter 4. Hydrokinetic turbines blade design

4.4.6. Effect of blade shape on Cp


Blades’ shape affects turbines’ efficiency. In Figure 4.8, Cp for three different types of
wind turbine blade shapes is compared. According to this figure, about 7% higher in Cp
is achievable using blade shape close to the theoretical optimal shape C compared to the
simplified rectangular shape A. This can be a significant loss, especially for a
hydrokinetic turbine operating in very low speed current in a small river where the total
available power is limited. An optimum blades shape may be thought to be more
complex, expensive and difficult to construct, but this research has demonstrated in
section 4.6 that the design construction of a fairly optimum shape blades can be
relatively easy.

Figure 4.8. The effect of blade’s shapes on rotor power Cp [4.15]

Water current energy for remote community: Design and testing of a clog-free horizontal axis 98
hydrokinetic turbine
Chapter 4. Hydrokinetic turbines blade design

4.4.7. The effect of blade twist on Cp

Figure 4.9. The effect of blade twist on rotor Cp [4.15]

Figure 4.9 shows that about 5 % more power is gained with the twisted blade A
compared to the untwisted blade C. A simpler linearly twisted blade B performs almost
similar to twisted blade A, suggesting that simpler twist can be adopted without
significant loss but as stated earlier, optimum blades featuring twist, varied blade
thickness and tapper are not too hard to construct once a blade template has been made.
So a hydrokinetic turbine prototype developed in this research is equipped with blades
close to optimum shape and twist.

4.4.8. Blade effective area


Figure 4.10 shows that the outer and more slender half of the blades is where most of its
power is produced. Because of this, the outer part of the blade must be properly
designed. Highly efficient airfoils or hydrofoils built as close as possible to the

Water current energy for remote community: Design and testing of a clog-free horizontal axis 99
hydrokinetic turbine
Chapter 4. Hydrokinetic turbines blade design

theoretical optimal shape should be used. In reference to Figure 4.10, only about 9% of
power is lost if about a third (35%) of the blade (from the root) is omitted. This small
loss in power has become the basis of the rotor design used in this research, where about
a third of the blade length is replaced with pivoting metal spokes covered by a large
nose cone for debris and impact management purpose.

Figure 4.10. Cp performances when parts of the blade are omitted [4.15]

Water current energy for remote community: Design and testing of a clog-free horizontal axis 100
hydrokinetic turbine
Chapter 4. Hydrokinetic turbines blade design

4.4.9. Hydrofoil terminologies


The terminologies used to describe a hydrofoil are the same as for an airfoil. Figure 4.11
shows terminologies use to describe a turbine blade [4.20]. The straight line connecting
the leading edge and the trailing edge of the foil is term as the chord line, c. This is the
parameter that must be calculated in order to get blades’ coordinates that represent its
shape and thickness of a blade and this is demonstrated in the thesis. A point called the
hydrodynamic centre is located along the chord line, a quarter of c from the leading
edge. This is the point where hydrodynamic forces intersect, logically the centre of the
blade spars. The median line or the camber line is an arc line dividing the foil in the
middle equally and the distance between this line and the chord line shows the degree of
camber an air or hydrofoil has.

Figure 4.11. Hydrofoil terminologies [4.20]

Water current energy for remote community: Design and testing of a clog-free horizontal axis 101
hydrokinetic turbine
Chapter 4. Hydrokinetic turbines blade design

4.4.10. Operation of a hydrofoil


The principle of hydrofoil operation lies in the pressure difference when water passes
over the “upper” (convex or suction) side and the “lower” (concave or high pressure)
side of the hydrofoil. Water current incident on the blades turns the blades at a certain
speed, resulting in a relative current coming at an angle under and near the nose (leading
edge) of the blade as shown in Figure 4.12. Water current passing on the upper side of
the blade travels faster than that below causing a pressure drop, resulting in a lift force
perpendicular to the relative current direction. This force further creates a driving force
that turns the blades faster in its rotation plane.

Figure 4.12. Water current flow around a hydrofoil and forces created

If the blade turns too fast, local pressure on blade drops below vapor pressure resulting
in small vapor cavities or bubbles to build up at the affected area and cause water flow
to separate from the surface of blade as illustrated in Figure 4.13. This phenomenon
called cavitation is considered as one of the most extreme problem faced by
hydrokinetic turbines because of its ability to damage turbines’ blades and causes
reduction in efficiency [4.4]. Beside losing power due to flow separation, the shock
pressures created by continuous forming and collapsing of bubbles during cavitation is

Water current energy for remote community: Design and testing of a clog-free horizontal axis 102
hydrokinetic turbine
Chapter 4. Hydrokinetic turbines blade design

Figure 4.13. Cavitation on blades at high speed operation

Figure 4.14. Blade damaged by cavitation [4.21]

known to engrave and deformed blade surfaces, therefore causes further reduction
blade’s efficiency. Cavitation problem is also a common problem faced by pumps

Water current energy for remote community: Design and testing of a clog-free horizontal axis 103
hydrokinetic turbine
Chapter 4. Hydrokinetic turbines blade design

impellers and boats or ships propellers [4.4]. An example of the effect of cavitation on
propellers is shown in Figure 4.14 [4.21].

Cavitation inception can be predicted using equation (4.1) below [4.22]. A


dimensionless number call cavitation number denoted by σ is defined as:


(4.1)

Where PAT = atmospheric pressure (101320N/m2),


⍴ = water density (1000 kg/m3 for freshwater, 1025 kg/m3 for seawater)
H = rotor depth of immersion,
PV = vapor pressure (2000 N/m2) and
V = velocity at blade tip ( ), where U is water current velocity,
Ω is rotor rotational speed rad/s and R blade tip radius [4.16].

The higher the cavitation number is, the less possibility cavitation inception to occur
and lower number will increase the chance for cavitation inception to occur. Cavitation
study found in [4.16] shows that a turbine cavitation starts to appear at σ = 0.89. Figure
4.15 shows a test turbine operate normally at σ = 1.2 but cavitation appeared at σ = 0.64
as shown in Figure 4.16.

Equation 4.1 indicates that cavitation number is a function of pressure, depth of


immersion and blade tip speed, however tip speed and depth of immersion are two
parameters that can control by operators during operation. Rotor immersion depth must
be sufficient so that ventilation (sucking of air by blades) and cavitation do not occur
easily. For example, 1.5 m rotor immersion was used in [4.18]. Therefore, operating at a
correct tip speed or tip speed ratio is important for a hydrokinetic turbine to operate
effectively. As stated earlier, blade tip speed is normally limited to not more that 12 m/s
or tip speed around 3 in order to avoid cavitation.

Water current energy for remote community: Design and testing of a clog-free horizontal axis 104
hydrokinetic turbine
Chapter 4. Hydrokinetic turbines blade design

Figure 4.15. Normal operation at σ = 1.2

Figure 4.16. Cavitation inception at σ = 0.64

4.4.11. Hydrodynamic forces


Figure 4.17 (adapted from wind turbine, reference [4.14]) shows the direction of forces
resulting from the interaction of the incoming water current and the rotation of the
turbine blade. This action results in a relative current faster than the original current.
The angle between the chord line and the direction of the relative current felt by the

Water current energy for remote community: Design and testing of a clog-free horizontal axis 105
hydrokinetic turbine
Chapter 4. Hydrokinetic turbines blade design

blade is represented by α and this angle is called the angle of attack. This angle has a
significant influence on the overall performance of the rotor.

Figure 4.17. Forces created by the interaction of water current and blade rotation [4.12] (adapted)

The relative current produces a lift force at the hydrodynamic centre perpendicular to
the direction of the relative current and a drag force in the direction of the relative
current. The vector sum of the lift and drag forces creates a resultant force. It is this
resultant force that further gives rise to the required driving force to rotate the blades.

Besides producing the needed driving force, the resultant force also creates a large
thrust force which intensifies significantly if the rotor turns at high speed. In fact,
managing this force on site is quite challenging because it tends to bend the turbine’s
leg (structure) or swing the rotor out of the water. The force also tends to lift the
downstream part of turbines pontoons up in the air and try to submerge the front part. In
one installation shown in Figure 2.19 (b) in Chapter 2, sand bags were used to level the
pontoon during operation. Therefore the understanding of all these hydrodynamic forces
is essential in the development of a turbine prototype developed in this research.

Water current energy for remote community: Design and testing of a clog-free horizontal axis 106
hydrokinetic turbine
Chapter 4. Hydrokinetic turbines blade design

4.4.12. Blade design equations


Blade element momentum theory or BEM theory developed for wind turbines is used to
design hydrofoils blades or optimum blades for a hydrokinetic turbine prototype in this
research, following verification and validation stated in section 4.2. According to [4.6],
BEM theory is a combination of two theories namely conservation of momentum theory
and blade element theory. Conservation of momentum theory states that loss of pressure
or momentum through the rotor plane is caused by work done on the turbine blades by
the moving fluid whereas blade element theory is the analysis of forces acting
independently on individual blade elements and sum them up to give total forces and
moments exerted on the turbine. The combination of these two theories result in
equations (4.2), (4.3) and (4.4) which are used in this research to determine blade cord
lengths C and twist angles β for every element or section at any distances along the span
of a blade.

The blade design equations are based on Figure 4.18, which is the same as Figure 4.17
but with vector forces removed. This figure contains velocity vectors which become the
basis of equations 4.2 to 4.4 and the explanation on how the equations are derived can
be found in [4.12] and examples of their application in wind turbine blade design, which
can be adopted for hydrokinetic turbine blade design, can be found in [4.27].

(4.2)

(4.3)

(4.4)
Where = chord length (mm),
= radius or distance from axis of rotation of a section on the blade (mm),
= number of blades,
Cld = design lift coefficient,
= angle of relative water current velocity to direction of blade travel (deg.),

Water current energy for remote community: Design and testing of a clog-free horizontal axis 107
hydrokinetic turbine
Chapter 4. Hydrokinetic turbines blade design

= twist angle (deg.),


αd = design angle of attack (deg.) and
λd = design tip speed ratio (tsr).

Figure 4.18. Velocity vector diagram for deriving blade design equations [4.12] (adapted)

The blade design equations are used to determine an optimum rotor with wake rotation
consideration. Wake rotation refers to the rotation ω of air flow behind the rotor of a
wind turbine (assume the same in hydrokinetic turbine) which is in the opposite
direction to the rotor during operation. This phenomenon can be related to the torque
produced by the rotor, so the vector representing the relative velocity of air or water
hitting the blade is equal to Ωr + 1/2ωr, where Ω is the blade angular velocity and ω is
the wake rotation angular velocity, as shown in Figure 4.18.

Water current energy for remote community: Design and testing of a clog-free horizontal axis 108
hydrokinetic turbine
Chapter 4. Hydrokinetic turbines blade design

Other vectors include the relative water current velocity represented by Vrel, water
velocity at the blade is represented by (1-a)V where “a” denotes the induction factor
which accounts for slowing down of the water as kinetic energy is extracted by the
blades and its value is ideally 1/3 of the water velocity V. The Angle φ in Figure 4.18
represents the angle of relative water current, ß represents the twist angle and α
represents the design angle of attack.

While turbine blades can be designed using CAD software and constructed
automatically using CNC machines, making wooden blades manually require a template
to be made as close as possible to the theoretical optimum shape. In order to do this, the
optimum hydrofoil’s profile (width/chord, twist and thickness) at some known distances
along the whole length of the blade must be obtained. The chord lengths or width and
twist angle of a blade section at a specific distance from the root can be obtained using
equation 4.2 and equation 4.3 respectively. However, before these equations can be
used, a third equation (equation 4.2) must be solved first.

4.5. Hydrokinetic turbines blade design


In order to design an optimum turbine blade, the blade length or rotor radius must be
calculated. The blade length is related to how much power is to be produced, so using
the power equation discussed in section 4.4.2, the blade length which is equivalent to
the rotor radius can be calculated. The blade length is then divided into some equal
sections and the profiles’ coordinates and the twist angle for each section are obtained
using BEM theory equations. Solidworks software imports these coordinates to produce
a 3D model of the blade and also to print section profiles in 1:1 scale for blade template
making. The following sections explain blade design procedures starting from rotor
sizing down to calculating blade profile coordinates.

4.5.1. Rotor sizing


Turbine rotor sizing in this research starts by first estimating the power required by a
remote home. Energy requirement in remote villages should not be too demanding as it
is in urban areas. For example, alternating current (AC) in the range of 200 watts is
considered sufficient for a single remote home [4.23]. Power in this magnitude, using
conventional 240 volt, 50 Hertz could enable the use of a mini refrigerator and adequate
lighting (using energy saver light globes). After the determining the amount of power

Water current energy for remote community: Design and testing of a clog-free horizontal axis 109
hydrokinetic turbine
Chapter 4. Hydrokinetic turbines blade design

required, the length or the radius of turbine blades is calculated using equation 2.1
found in Chapter 2, repeated below.

Where P = power (Watt),


 = density of water (kg/m3) = 1000 kg/m3 for fresh water,
A = cross sectional area (m2),
V = velocity of water current (m/s),
Cp = coefficient of power or turbine efficiency and
η = combined drive train & generator efficiency.

Before the length of the blade can be calculated, the design water current velocity V, the
coefficient of power Cp and the combine drive train and generator efficiency η must be
determined first. In this research, the target water current velocity is 1 m/s because
water current in this range is quite common in remote areas in Sarawak, Malaysia.
Assume that a fairly efficient turbine having Cp of 35% is used even though a small
turbine reported in [4.16] can achieve Cp of 46%. Assume also the combined drive train
and generator efficiency is equal to 60%. Both the Cp and η are quite conservative
assumptions. This results in and overall efficiency or water to wire efficiency (Cp*η) of
21%, a typical value for a small horizontal axis hydrokinetic turbine for electric
generation as reported in [4.24]. The radius and hence the length of the turbine blades
can then be calculated using equation 2.1 as shown below.

Water current energy for remote community: Design and testing of a clog-free horizontal axis 110
hydrokinetic turbine
Chapter 4. Hydrokinetic turbines blade design

From the calculation above, the radius of the turbine rotor (from the centre point of the
rotor) is equal to 0.779 meter but for simplicity, a blade length of 0.8 meter is used.
However, the final blades are shorter than this because part of the root will be replaced
by the radius of the rotor hub used. So, this calculation reveals that a hydrokinetic
turbine rotor having radius of about 0.8 m or 1.6 m diameter is required to generate 200
of electric power in 1 m /s water current. But if the turbine operates in higher water
current velocity, say 2 m/s, the turbine rotor will decrease significantly to only about 0.5
m diameter.

4.5.2. Optimum blade profile


Generally, an optimum turbine blade (wind or hydrokinetic) is narrower at the tip and
wider near the root. The blade is also thicker at the root, where greater flexural strength
is required, and gradually becomes thinner towards the tip where drag must be
minimized. The twist of the blade is significantly greater at the root compared to the tip.
These geometries are important in blades design and their effects on turbine
performance have been explained in section 4.4.1 to 4.4.8. Some turbine blades use
more than one airfoil or hydrofoil profile in order to attain maximum performance in
term of efficiency and strength. For example, a high performance wind turbine blade
used in DEBRA 25 turbine, shown in Figure 4.19 featured six different types of airfoils
[4.20]. However, hydrokinetic turbines use fewer number of hydrofoils types. A small
diameter horizontal axis hydrokinetic turbine tested in [4.16], used five hydrofoils
comprising of NACA 63-812, 63-812, 63-818, 63-821 and 638-24. Another horizontal
axis hydrokinetic turbine tested in [4.18] used a couple of hydrofoils where an ad hoc
thick hydrofoil was used between 15% and 35% of the blade radius for strength and a
thinner GT1, modified from S805 was used for the rest of the blade for high
hydrodynamic efficiency. However, one example of hydrokinetic turbine tested in
[4.28] used only one profile that is the NACA 002.

For the purpose of this research, a single S822 airfoil (used as hydrofoil) was adopted so
as to simplify wooden blade making in remote communities’ villages. The S822 is
chosen because of its resistant to fouling. According to [4.26], the S822 performed
better when dirty compared to the NACA airfoils. Even though it is slightly less
efficient, the fouling resistant capability is regarded more important especially when

Water current energy for remote community: Design and testing of a clog-free horizontal axis 111
hydrokinetic turbine
Chapter 4. Hydrokinetic turbines blade design

used in hydrokinetic turbine blades where water currents carry debris, dirt, tiny plant
and organism which could foul turbine blades easily.

Figure 4.19. Turbine blade comprising of multiple airfoils [4.20]

4.5.3. Determining blade coordinates


The coordinates of airfoils or hydrofoils are normally given as a function of chord
length C. For example the coordinates for the S822 used in this work are given in the
form of x/c, y/c. These coordinates represent the profile (shape and thickness) of a blade
at a particular section or blade element r distant from the rotor centre along the blade
span and this can be worked out using equation (4.2). The twist angles for each
section at any distant r can be found using equation (4.3). Since both equations depend
on (angle of relative water current to the plane of rotation), an intermediate equation
(4.4) is required to obtain the chord lengths and the twist angles of each blade section.

Water current energy for remote community: Design and testing of a clog-free horizontal axis 112
hydrokinetic turbine
Chapter 4. Hydrokinetic turbines blade design

4.5.3.1. Calculating chord length C

The chord length C of a blade section at any distant r along the blade radius can be
found using equation (4.2). This equation requires and to be predetermined. As
stated in section 4.4.3, the number of blades used in this research is three ( ) and
mainly for higher efficiency and smoother output torque.

The lift coefficient for selected airfoils (use as hydrofoils) can be obtained from wind
tunnel test data. In this research, the S822 airfoil (as hydrofoil) is adopted on the basis
explained in section 4.5.2. However, there is no report about this particular aerofoil
being used as a hydrofoil for hydrokinetic turbine so far. Nevertheless, airfoils test data
tested at various Reynolds numbers are usually used in hydrokinetic turbines’ blade
design so thus the data for the S822 airfoil.

According to [4.6], the net result of water having high density but low velocity and air
having low density but high stream velocity is that the Reynolds numbers used for wind
turbines are in the same range with those for hydrokinetic turbines. Therefore the same
experimental airfoil data can be used in hydrokinetic turbines. Reynolds number is a
vital dimensionless parameter in fluid dynamic [4.27]. It is defined as;

(4.5)

Where V = water current speed,

C = chord length and

= kinematic viscosity of water ( ).

In this research, the target water current speed is 1 m/s and the chosen tip speed ratio is
4. These result in the speed of water current velocity V at the tip of the blade equal to 4
m/s. And according to Table 4.2, the tip chord length of the blade designed in this
research is 171 cm or 0.171 m. Therefore using equation (4.5), the Reynolds number
for the blade tip can be calculated as below;

Water current energy for remote community: Design and testing of a clog-free horizontal axis 113
hydrokinetic turbine
Chapter 4. Hydrokinetic turbines blade design

Figure 4.20 The S822 lift and drag plots [4.19]

The calculation above shows that the Re calculated for the turbine blade falls in about
the range of Re used in wind turbines. In reference to Figure 4.20, the test data for S822
airfoil are plotted for Re = 100,000 to 500,000. Since there is no plot for Re = 600,000,
S822 data for Re = 500,000 is used.

4.5.3.2. Finding maximum nd

The highest value for coefficient of lift is obtained by drawing a straight line, tangent
to the plot for Re = 500,000 as shown in Figure 4.20(a). Drawing a horizontal
line from the point where the line touches plot gives equal to 0.8. The angle of

Water current energy for remote community: Design and testing of a clog-free horizontal axis 114
hydrokinetic turbine
Chapter 4. Hydrokinetic turbines blade design

attack , a variable required in equation (4.3) can then be found by plotting a horizontal
line from value of 0.8 until it touches the plot for Re = 500,000 as shown in
Figure 4.20(b). Therefore the corresponding value of (which in this case is 5 degrees)
o
is the required angle of attack, therefore =5 .

The relative angle of water current velocity φ can be worked out using equation (4.4)
but the design tip speed ratio tsr or has to be determined first. As mentioned earlier,
choosing the right tsr is important in order to avoid power lost and blade damage by
cavitation. This is important because the blades of a hydrokinetic turbine are perhaps
the most delicate and sensitive part of the turbine. Furthermore, replacing turbine blades
on site can be an expensive, dangerous and difficult exercise which will result in
prolonged power cutoff.

In this work, a tsr = 4 is chosen in order to be consistent with statement stated in


section 4.4.3 where the tip speed of turbine blades must be kept below 12 m/s or tsr
below 7. For design water current velocity of 1 m/s water current, the tip speed is only
4 m/s. Even though during wet season, water current velocity in mid-course of some
rivers in Sarawak, Malaysia may increase to 1.5 m/s but the tip speed is still within a
safe value of 6 m/s.

Now that the values for B and have been determined, and have been found, the
value for and can now be calculated for any value of along the blade span. For
simplicity, the whole blade length (which is 0.8m) is divided into 10 equal segments so
adjacent sections are separated by an 80 mm distance (0.8/10 = 0.08 m or 80mm). The
sections are denoted by , to , where is the section nearest to the root and is
at the tip of the blade. In this work, an Excel spreadsheet shown in Table 4.2 was used
to obtain the values of and for the value of to .

The second last column of Table 4.1 are the value for chord length , calculated using
equation 4.2 for to along the blade span. Notice that the value of increases
from the tip towards the root except for station 1 or . On the hand the last column of
Table 4.3 are the twist angle calculated using equation (4.3) for to . Notice also
that the value for are larger near the root compared to the tip. The patterns of the
and values shown in Table 4.1 can be seen to resemble basic shape of optimum blades
explained in sections 4.4.4, 4.4.7 and 4.4.8.

Water current energy for remote community: Design and testing of a clog-free horizontal axis 115
hydrokinetic turbine
Chapter 4. Hydrokinetic turbines blade design

Table 4.2. An Excel spreadsheet for determining sections chord length and twist angle

4.5.3.3. Calculating blade profile coordinates

After the value of at every station has been found, the next step is to multiply these
values by the non-dimensional coordinates of the S822 in Table 4.3 [4.12]. Notice that
the coordinates are given in the form of , In order to obtain and , the
coordinate values for , for the S822 airfoil (used as hydrofoil) are multiplied
the values of found in Table 4.1 for to . For example, the value for at or
station 10 which is located at the tip of the blade is 0.171 meter (see Table 4.2) and the
value of the first set of coordinate (upper side) in Table 4.3 are 0.00012 mm,
0.00132 mm. Multiplying these two set of numbers by the value of C at will
produce the first set of coordinates 0.00002 mm, 0.00023 mm. This procedure
is repeated for all values of , (upper and lower surface coordinates) given in
Table 4.3 in order to obtain a 2D coordinate of the blade profile at station .

The blade profile coordinates for other stations can be obtained using the same method
as . To ease and to speed up this process, another Excel spreadsheet can be used.
After all coordinates at all sections are calculated, they can then be scatter plotted and
printed in 1:1 scale while in Excel environment. The scaled print can then be used to
form 3D or physical templates for each station as demonstrated later in section 4.4.6.
The Excel scatter plots for (near root) and (at the tip) obtained in this work are
shown Figures 4.21(a) and 4.21(b) respectively.

The coordinates obtained for to can also be exported to CAD (computer aided
design) software for precise to scale printings, modeling and simulation. In this work,

Water current energy for remote community: Design and testing of a clog-free horizontal axis 116
hydrokinetic turbine
Chapter 4. Hydrokinetic turbines blade design

the blade profile coordinates are exported to Solidworks software to form a blade model
for visualization and demonstration purpose. For Solidworks to accept the coordinates, a

Table 4.3. The coordinates of the S822 airfoil [4.12]

Figure 4. 21(a) Profile plot at r2 (near the root), (b) Profile plot at section r10 (at the tip)

Water current energy for remote community: Design and testing of a clog-free horizontal axis 117
hydrokinetic turbine
Chapter 4. Hydrokinetic turbines blade design

third coordinate is required. For example the first coordinate for become
0.03348 mm, 0.36828 mm, 0 or simply 0.03348, 0.36828, 0. Each S822 section
profile is comprised of 33 sets of coordinates for the upper side and 32 sets for the lower
side profile. Therefore a total of 65 sets of coordinates are required to form complete 2D
section profile for S822 airfoil as hydrofoil.

However, the x, y, z coordinates in Excel format are not recognized by Solidworks.


Therefore all coordinates for section profiles must be converted to NotePad format by
simply copying Excel table content and pasting them in a new NotePad document.
Each section from to will have one NotePad document containing 65 sets of
coordinates and will be exported to Solidworks one file at a time.

While in sketch mode of Solidworks software, the Arch function can be activated to
import the NotePad files containing blade profile coordinates for each section one by
one. For example, the set of NotePad coordinates x, y, z for is imported first using
the Arch function in Solidworks and the software will automatically sketch the profile
for this particular section on the computer screen. Before a NotePad file containing 65
sets of coordinates representing the next section profile is imported, a new plane is
inserted 8 cm in front of this first profile sketch. This procedure is repeated for all the
remaining sections.

After all the section profiles have been sketched, their corresponding twist angles
tabulated in the last column of Table 4.2 are then added to the section profile sketch
using the Rotate command found in Move Entity menu of Solidworks. For example,
using Rotate command, a twist angle of 7° is added to the first profile sketch of .
Upon execution of the 7° Rotate command, the sketch will rotate to show 7° twist
angle. This procedure is again repeated for the rest of the sections and after all the
profile sketches at all sections are given their appropriate twist, the entire sketch should
look like the image in Figure 4.22(a).

Water current energy for remote community: Design and testing of a clog-free horizontal axis 118
hydrokinetic turbine
Chapter 4. Hydrokinetic turbines blade design

(a) (b) (c)

Figure 4.22. Blade drawing using Solidworks software

At this stage, the sketch shows blade profiles or contours at every section and their
twists angles. In order to get a 3D model, the Loft command is used. While in Features
mode of the Solidworks software, the Loft command is selected. This command
requires all section profile sketches to be is selected and upon executing the Loft
command, all sections profiles in Figure 4.23(a) will be lofted to give the image shown
in Figure 4.22(b). Notice that the images in Figure 4.22(a) and (b) clearly show that
blade shape design using BEM theory and modeled using Solidworks is thicker near the
root and becoming thinner towards the tip. The contours also show that the blade is
generally narrower at the tip and wider toward the root, and becomes narrower again at
the root. The twist angles are clearly increasing toward the root. The shape obtained is
about the same as the optimum blade shape described earlier.

Depending on the radius of the turbine hub, the last two section profile sketches may be
omitted as shown in Figure 4.22(c). This final image can be used in further development
of a turbine system. For example, the image can be assembled or integrated with other
parts of the turbine to form a complete model in Solidworks. Computer codes suitable
for a CNC machine can also be generated from this image so that hydrokinetic turbine
blades can be milled from aluminum or other suitable materials automatically using a
CNC milling machine. However, this is not attempted in this work, in favor of simple
and more appropriate methods that are suitable for remote communities as explained in
the following sections.

Water current energy for remote community: Design and testing of a clog-free horizontal axis 119
hydrokinetic turbine
Chapter 4. Hydrokinetic turbines blade design

4.6. Optimum wooden blades construction


In order to be appropriate, wooden blades are preferred in this work. The main reason
for doing so is because wood is readily available in remote villages in Sarawak
Malaysia. Furthermore propellers made from wood have been proven successful in
small and large aircraft in the past. At least one hydrokinetic turbine company claims
that wood has good properties for underwater application and is now developing
wooden blades for their turbine [4.29].

A particular type of hard wood called “belian” is very well known to the remote people
of Sarawak for its strength, resistance to wood bugs and super durability even when
submerged in water. Local people use it for house pillars and the power utilities in
Sarawak use it for electric power transmission poles. This hard wood can last up to fifty
years or more. Some say that it lasts longer under water or underground. Therefore the
author plans to use this wood as hydrokinetic turbine blades, but for a demonstration
purpose, treated pine was used instead.

Instead of using a CNC machine, a blade template was made manually using common
wood working tools. The contours of every section were printed to scale using
Solidworks software (scatter plot in Excel can also be printed in the same manner), the
outlines were cut and traced on a piece of soft wood (pine was used) about 10 mm thick,
shown in Figure 4.23(a). The outlines on the wood were then cut carefully using a
coping saw as shown in Figure 4.23(b) and smoothed using a suitable file (a metal file
was used). Figure 4.23(c) shows all profile templates at all sections arranged in order
minus the twist angles. Notice that the chord increases significantly from tip towards
root but decreases slightly at the root.

Water current energy for remote community: Design and testing of a clog-free horizontal axis 120
hydrokinetic turbine
Chapter 4. Hydrokinetic turbines blade design

Figure 4.23. (a) to (c) Making section templates, (d) to (f) Making whole blade template

In order to construct a solid blade template, correct size of “filler” wood pieces were
used to fill the gaps between profile templates. The wood pieces used were 70 mm thick
(allowing for section profile templates which are 10 mm thick). They were cut slightly
longer than the chord length of the two section profiles in which the filler was intended
to be. Starting from the tip, section template was placed on one side of the filler
wood, given a 7 degrees twist and the final position was traced on the filler wood. On
the opposite side of the same piece of wood, section template with 10 degrees twist
angle is traced. The filler wood bearing the contours and twist angles of on one side
and on the other was then cut using a band saw, about 3 mm to 5 mm away from the
contour lines as shown in Figure 4.23(d). This process was repeated for the rest of the
sections.

After all the filler wood pieces have been cut, they were further shaped and smoothed
using a wood file or rasp. Then they were arranged in correct order and in such a way
that they were sandwiched in between their corresponding section templates as shown
in Figure 4.23 (e).

Two screw holes were prepared on each section template and shaped filler wood.
Starting from the lower most usable section template (in this case ), wood glue was
applied sparingly on both sides. The corresponding shaped filler was stacked on top and

Water current energy for remote community: Design and testing of a clog-free horizontal axis 121
hydrokinetic turbine
Chapter 4. Hydrokinetic turbines blade design

two long screws were driven down from the first filler, through the first section template
and to the base of the blade template as shown in Figure 4.23(e). The screws were used
to clamp the filler and the section templates while the glue was curing. In order to allow
the next template to be mounted on top, the screw heads were countersunk a little below
the filler surface. The rest of the templates and fillers were mounted in the same manner
one on top of the other and left to dry for at least 24 hours. Using one side of each
section template as guides, the filler wood which was slightly thicker than the section
template was grounded or flushed carefully using an angle grinder fitted with a medium
grit flap disc. The blade template was further shaped and smoothed using cyclic power
sanders and sanded manually using fine sandpaper. The result is shown in Figure
4.23(f) and was used as a blade template for making three identical optimum wooden
blades to make up a three-bladed hydrokinetic turbine rotor. In order to assist identical
blade construction, two types of blade copy jigs were constructed using of-the-shelf
components and materials. The use of these jigs for making identical blades is explained
in Appendix 5.

4.6.1. Issues in wood based hydrokinetic turbine blades.

The wooden blades designed and constructed for a hydrokinetic turbine prototype in
this research are almost entirely made by hands and of course with the help of common
tools and homemade jigs. Even though the design and construction processes, including
selecting wood for making identical blades were carried out carefully, the blades
produced are still not as precise as the commercial blades professionally produce for
wind turbines. Blade making errors and different in wood density could cause
imbalance when turbines’ rotor rotate at high speed. However, error investigation on
hand-carved blades similar to report in [4.30] was not contemplated and measurement
of mass and centre of mass as in [4.31] were not carried out mainly because blade
precision is fairly emphasized in this research. Furthermore this work is in its early
stage where a set of fairly efficient blades are required in order to quickly test a clog
free rotor design ideas in order to prove that hydrokinetic turbine can be practical if use
in debris laden rivers in Sarawak, Malaysia.

In addition, it was assumed in this research that imbalance problem due to blade error
can be tolerated, especially in a hydrokinetic turbine prototype operating at low rpm in
low velocity water current. Water being much denser compared to air should be readily
to absorb vibration due to blades imbalance. Fortunately this was the case during the

Water current energy for remote community: Design and testing of a clog-free horizontal axis 122
hydrokinetic turbine
Chapter 4. Hydrokinetic turbines blade design

field test of a prototype developed in this research. Vibration due to blades errors and
other imbalance cause by other components making up the rotor was not a big issue
throughout the testing.

4.7. Conclusion
This chapter has explained that wind turbines’ design principles are valid for
hydrokinetic turbine provided that cavitation is avoided. Evidences of successful
applications of hydrokinetic turbine design tools based on wind turbines’ BEM theory
are stated. Upon confirming that wind turbine principles are valid for hydrokinetic
turbines design, the chapter proceeds to apply and demonstrate how these principles are
applied in the design of a hydrokinetic turbine prototype. A blade for a 200 W
horizontal axis hydrokinetic turbine has been designed successfully in this work. The
procedures involved predetermining blade features based on wind turbines’ design
practices, calculating blade section profile coordinates and twist angle using BEM
theory equations with the help of Microsoft Excel. The coordinate in Microsoft Excel
were converted to NotePad in order to be recognized by Solidworks software. The
section profiles coordinates are imported into Solidworks to produce 1:1 scale printout
for manual blade template making. The steps in wooden blade template making were
explained and copying this template to make three identical blades for a rotor using two
type of jigs is included in Appendix 5. 3D models and CNC code for blade making
using CNC mill can also be produced using Solidworks but was not attempted, in favor
of appropriate methods for remote communities to comprehend.

References
[4.1] Mukherji, S., et al. Numerical investigation and evaluation of optimum
hydrodynamic performance of a horizontal axis hydrokinetic turbine. Renewable
and Sustainable Energy, 2011. 3(063105(2011))

[4.2] Wang, D., Atlar, M. and Samsom, R. An experimental investigation on


cavitation, noise, and slipstream characteristics of ocean stream turbines. Power
and Energy, 2006. 221(Special): p. 219 -231.

Water current energy for remote community: Design and testing of a clog-free horizontal axis 123
hydrokinetic turbine
Chapter 4. Hydrokinetic turbines blade design

[4.3] Khan, M.J., Iqbal, M.T. and Quaicoe, J.E. River current energy conversion
systems: Progress, prospects and challenges. Renewable and sustainable Energy
Reviews, 2008. 12(2008): p. 2177-2193.

[4.4] Bahaj, A.S. and Myers, L.E. Fundamentals applicable to the utilization of
marine current turbines for energy production. Renewable Energy, 2003.
28(2003): p. 2205-2211.

[4.5] Gounder, J.N., Ahmed, M.R. and Lee, Y.H. Numerical and experimental studies
on hydrofoils for marine current turbines. Renewable Energy, 2012. 42(2012)

: p. 173-179.

[4.6] Sale, D. Hydrodynamic optimization method and design code for stall-regulated
hydrokinetic Turbine rotors. ASME 28th International Conference on Ocean,
offshore, and Arctic Engineering 2009: Honulu, Hawaii.

[4.7] Bahaj, A.S., Batten, W.M.J. and McCann, G. Experimental verifications of


numerical predictions for the hydrodynamic performance of horizontal axis
marine current turbines. Renewable Energy, 2007. 32(2007): p. 2479-2490.

[4.8] Marsh, G. Tidal turbines harness the power of the sea. Reinforced plastics,
2004. 48(6): p. 44-47.

[4.9] Rahman, H. Turbine may carpet more of E. River floor. 2009 [cited 18 May
2009]; Available from:
http://www.citylimits.org/content/articles/viewarticle.cfm?article_id=3702

[4.10] Tyler, R.N. River Debris: Causes, Impacts, and Mitigation Techniques. 2011
[cited 20 Nov. 2012]; Available from:
http://www.uaf.edu/files/acep/2011_4_13_AHERC-River-Debris-Report.pdf.

[4.11] Fraenkel, P.L. Marine current turbines: An emerging technology. 2004 [cited 20
Nov. 2012]; Available from:
http://www.ifremer.fr/dtmsi/colloques/seatech04/mp/proceedings_pdf/article_ab
stract/4.%20courants%20marins/4.2.MCT.pdf.

[4.12] Manwell, J.F., Mcgowen, J.G. and Rogers, A.L. Wind energy explained theory,
design and application. 2nd Ed. 2009. West Sussex, UK: John Wiley & Sons.

Water current energy for remote community: Design and testing of a clog-free horizontal axis 124
hydrokinetic turbine
Chapter 4. Hydrokinetic turbines blade design

[4.13] Orme, J.A.C., Masters, I. and Griffiths, R.T. Investigation of the effect of
biofouling on the efficiency of marine current turbines. In MAREC 2001.
France.

[4.14] Kirke, B.K. and Lazaukas, L. Variable pitch Darrieus water current turbines.
Fluid Science and Technology, 2008. 3(3): p. 430-438.

[4.15] Hau, E. Wind Turbines: Fundamentals, Technologies, Application, Economics.


2000, Berlin: Springer.

[4.16] Bahaj, A.S., et. al. Power and thrust measurements of marine current turbines
under various hydrodynamic flow conditions in a cavitation tunnel and a towing
tank. Renewable Energy, 2007. 32(2007): p. 407 – 426

[4.17] Swenson, W.J., The evaluation of an axial flow, lift type turbine for harnessing
the kinetic energy in tidal flow, in Australian Universities Power Engineering
Conference AUPEC. 1999: Darwin, Australia.

[4.18] Coiro, D.P., et al., Horizontal axis tidal current turbine: Numerical and
experimental investigations, in OWEMES. 2006: Civitavecchia, Italy.

[4.19] Tiago, G.L. The state of art of hydrokinetic power in Brazil, in Waterpower
XIII. 2003: Buffalo, New York.

[4.20] Gasch, R. and Twele, J. Wind power plant fundamental, design, construction
and operation. 2002, Berlin: Solarpraxis.

[4.21] Cavitation propeller damage. [Cited 20 Nov 2012]. Available from:


http://upload.wikimedia.org/wikipedia/commons/e/e6/Cavitation_Propeller_Da
mage.JPG.

[4.22] Batten, W.M.J., et al. Hydrodynamics of marine current turbines. Renewable


Energy 2006. 31(2006): p. 249-256.

[4.23] Freere, P. Electronic Load/Excitation for a Self-excited Squirrel Cage Generator


Micro-hydro. In IEEE Fifth International Conference on Electrical Machines and
Drives. 1991.

Water current energy for remote community: Design and testing of a clog-free horizontal axis 125
hydrokinetic turbine
Chapter 4. Hydrokinetic turbines blade design

[4.24] Swenson, W.J., The evaluation of an axial flow, lift type turbine for harnessing
the kinetic energy in tidal flow. In Australian Universities Power Engineering
Conference AUPEC. 1999: Darwin, Australia.

[4.26] David, A.S., Wind turbine technology. 1998, New York: The American Society
of Engineers.

[4.27] Lysen, E.H. Introduction to wind energy. 2nd ed. 1983, The Netherlands: CWD.

[4.28] Lazauskas, L. and Kirke, B.K. Modeling passive variable pitch cross flow
hydrokinetic turbines to maximize performance and smooth
operation.Renewable Energy, 2012. 45(2012): p. 41-50.
[4.29] Abelson, A. and Lie E. Research on wooden tidal energy turbine blades. [cited
21 October 2011]; Available from:
http://www.renewableenergyworld.com/rea/news/print/article/2011/research-on-
wooden-tidal-energy-turbine-blades.

[4.30] Clausen et al. Estimating the braking stresses in a small wind turbine blade. In
International Workshop on Small Scale Wind Energy for Developing Countries.
2009: Nairobi, Kenya.

[4.31] Hitz, K. and Wood, D. On blade matching after batch production. In


International Workshop on Small Scale Wind Energy for Developing Countries.
2009: Nairobi, Kenya.

Water current energy for remote community: Design and testing of a clog-free horizontal axis 126
hydrokinetic turbine

You might also like