You are on page 1of 77

Anton de Kom Universiteit van Suriname

Institute for Graduate Studies and Research


Renewable Energy Technology

RENEWABLE ENERGY
TECHNOLOGY RET

Experimental and numerical investigation


on the hydrodynamic performance of the
HK-10 hydrokinetic turbine

Student: Supervisor:
Corrado Sirianni (16RET1012) dr. ir. N. R. Nannan

Reviewers:
Prof. dr. A. Guardone
dr. ir. R. Sidin

Date:
April 17, 2020

Kennis maken en kennis delen in duurzaam partnerschap


Corrado Sirianni (16RET1012)
Experimental and numerical investigation on the hydrodynamic performance of the HK-10 hydrokinetic
turbine

Supervisor: dr. ir. N. R. Nannan


Reviewers: Prof. dr. A. Guardone & dr. ir. R. Sidin

Anton de Kom University of Suriname


Institute for Graduate Studies and Research
Renewable Energy Technology
Leysweg 86, P.O.B. 9212 Paramaribo, Suriname

Keywords: BEM analysis, dynamic similitude, hydrokinetic turbine, scaled model, water flume, wind
tunnel
Anton de Kom University of Suriname

Institute for Graduate Studies and Research


Renewable Energy Technology

Experimental and numerical investigation


on the hydrodynamic performance of the
HK-10 hydrokinetic turbine

Corrado Sirianni

Supervisor: dr. ir. N. R. Nannan


Mechanical Engineering Department,
Anton de Kom University of Suriname

Reviewers: Prof. dr. A. Guardone


Department of Aerospace Science and Technology
Politecnico di Milano, Italy

dr. ir. R. Sidin


Mechanical Engineering Department,
Anton de Kom University of Suriname

December 10, 2019


iv
ACKNOWLEDGEMENTS
First of all, I wish to acknowledge the support and great love of my family, my partner, Eva;
my mother, Reina; my father, Francesco; and my deceased grandmother, Henriette. The last
couple of years have not been very easy, nevertheless, they have always pushed me towards
my goals and have patiently provided me with time to complete my studies. I also want to
express my gratitude to the consecutive heads of the Mechanical Engineering discipline of the
Anton de Kom University, where I am employed, Jimmy Narain and Sanjeevkoemar Bissesar,
for their continuous support and the space they have given me to study during work hours,
even exonerating me of some of my responsibilities. Moreover, I would like to recognize the
invaluable opportunities that my supervisor, Ryan Nannan, dean of the Faculty of Technology
and former head of the Mechanical Engineering discipline, has given me. He has always
believed in me and always did his best to support my academic career, for which I am very
grateful. Of course I would like to thank the reviewers of this thesis, Ryan Sidin and Alberto
Guardone for their critical and insightful feedback, which for sure helped me write a better
academic piece. Furthermore, I want to pay my special regards to Rudi van Els, who has
offered me this amazing opportunity to visit the University of Brasília, where I have partly
conducted my research project. During the time spend in Brasília he has facilitated the
progress of my thesis project and has introduced me to very warm-hearted people, all very
generous to help achieve my goals. Also his son, Pedro, was very helpful in showing me
around and even helped me with some 3D printing issues. Thanks! Of course I wish to express
my gratitude to Taygoara Oliveira and Antonio Brasil, respectively head of the Mechanical
Engineering department and dean of the Faculty of Technology of the University of Brasília,
at the time of my visit, for their guidance and the freedom they have given me to use their
research facilities. It is with immense gratitude that I acknowledge the support and help of
Rafael Mendes, a PhD student at the University of Brasília. He was always available to answer
my questions on the experimental method employed for testing the hydrokinetic turbines in
the wind tunnel at the Energy and Environment laboratory (LEA). I am also indebted to the
kind-hearted guys of the 3D printing laboratory, Gabriel, Thiago, João and Andrea, all very
helpful in fabricating the printed parts of the lab-scale turbine. It gives me great pleasure in
acknowledging the help and interest shown by the students Gabriel Santos and Christian,
during the wind tunnel experiments; it was really nice getting to know you! I also want to
recognize the assistance and hospitality of Henrique and Danilo at the manufacturing
laboratory in Gama. I am also grateful for the freedom they had given me to work in the
laboratory until late at night; what a time! Although not directly involved in my thesis project,
I want to thank Angelika Namdar for all her efforts in organizing the MSc programme.
Circumstances have not always been ideal, but I am convinced that she did everything in her
power to keep the programme on rails. My heartfelt thanks! Last but not least I would like to
express my gratitude to all my friends and colleagues who have supported me during my
study period, namely Ramsay Mac Donald, my office-mate with whom I often had technical
discussions about my research project.

Corrado Sirianni, 17th of April 2020

v
vi
LIST OF SYMBOLS

Standard nomenclature

Symbol Description Unit


a Axial induction factor -
a’ Angular (or tangential) induction factor -
A Area m2
B Flume width m
c Chord length m
𝐶𝐷 Hydrodynamic drag coefficient -
𝐶𝐿 Hydrodynamic lift coefficient -
𝐶𝑁 Normalized blade force in the normal direction -
𝐶𝑃 Power coefficient -
𝐶𝑇 Normalized blade force in the tangential direction -
d Rotor diameter m
D Hydrodynamic drag force N
E Flume specific energy m
F Tip-hub correction factor -
𝐹𝑟 Froude number -
g Gravitational constant m/s2
i Counter variable -
𝑘𝑖=1…5 Fitting parameters -
L Hydrodynamic lift force N
𝐿𝐸 Laminar flow entrance length m
m Scaling parameter -
𝑚̇ Mass flow kg/s
M Torque on annular control volume Nm
N Number of blade elements -
𝑁𝐵 Number of rotor blades -
𝑝 Fluid pressure Pa
𝑝𝑖=1…5 Fitting parameters -
P Mechanical power W
𝑃𝑁 Blade force in the normal direction N
𝑃𝑇 Blade force in the tangential direction N
𝑞 Flume discharge per unit width m3/(m∙s)
𝑄 Flume volume flow rate m3/s
r Distance from rotor center to blade element m
r* Normalized distance, r/R -
R Rotor radius m
Re Reynolds number -
Rh Rotor hub radius m
T Thrust force N
u Flow velocity m/s
𝑢∞ Far upstream flow velocity m/s
𝑢𝜃 Tangential velocity of flow at actuator disc m/s

vii
Standard nomenclature (continued)

Symbol Description Unit


w Induced flow velocity m/s
x Distance in flume flow direction m
x’ Normalized distance in flume flow direction -
x* Cartesian coordinate to describe transition shape m
y Flume flow depth m
z Flume width variable m
𝑦𝑐 Critical flow depth m
z* Cartesian coordinate to describe transition shape m
𝛼 Angle of attack °
𝜃 Rotor blade pitch angle °
𝜆 Tip-speed ratio -
𝜆𝑟 Local tip-speed ratio -
𝜇 Dynamic viscosity Pa∙s
𝜈 Kinematic viscosity m2/s
Π1,2,3 Dimensionless groups -
𝜌 Mass density kg/m3
𝜏 Shaft torque Nmm
𝜙 Flow entrance angle °
𝜔 Rotational velocity of turbine rad/s

Superscripts and subscripts

(super/sub)script Description
(1) Flags pre-transition property
(2) Flags pre-transition property
crit Critical state property
d Flow property on actuator disc
fric Friction
h Rotor hub
max Maximum
min Minimum
mod Property related to lab-scale model
nom Nominal value
opt Optimal value
prot Property related to prototype
rel Relative value
w Flow property in the wake

viii
LIST OF ACRONYMS AND ABBREVIATIONS

ABS Acrylonitrile butadiene styrene


ADC Analog-to-Digital Converter
AdeKUS Anton de Kom University of Suriname
AES Applied Energy Services
BEM Blade Element Momentum
CAD Computer Aided Design
CFD Computational Fluid Dynamics
CV Control Volume
DC Direct Current
EDF Electric Ducted Fan
ESC Electronic Speed Control
HK Hydrokinetic
MIT Massachusetts Institute of Technology
MPPT Maximum Power Point Tracking
NACA National Advisory Committee for Aeronautics
PET Polyethylene terephthalate
PID Proportional Integral Derivative
PLA Polylactic Acid
PVC Polyvinyl chloride
PWM Pulse Width Modulation
RPM Revolutions Per Minute
RTV Room Temperature Vulcanizing
SCADA Supervisory Control and Data Acquisition
TSR Tip-Speed Ratio
UFPA Federal University of Pará
UnB University of Brasília
VFD Variable Frequency Drive
VoF Volume of Fluid

ix
x
TABLE OF CONTENTS
ACKNOWLEDGEMENTS ............................................................................................................................... V
LIST OF SYMBOLS ........................................................................................................................................ VII
LIST OF ACRONYMS AND ABBREVIATIONS ....................................................................................... IX
1 INTRODUCTION...................................................................................................................................... 1
1.1 BACKGROUND INFORMATION .............................................................................................................. 1
1.2 PROBLEM DEFINITION & OBJECTIVES ................................................................................................... 2
1.3 STRUCTURE OF REPORT ......................................................................................................................... 3
2 RELEVANT THEORY ............................................................................................................................... 5
2.1 A BRIEF ON HYDROKINETIC TURBINES ................................................................................................. 5
2.2 RANKINE-FROUDE ACTUATOR DISC MODEL ........................................................................................ 6
2.3 EFFECTS OF ROTATION ........................................................................................................................ 10
2.4 OPTIMAL FLOW CONDITIONS ............................................................................................................. 12
2.5 BLADE ELEMENT MOMENTUM THEORY ............................................................................................ 13
3 WATER FLUME TESTING FACILITY ................................................................................................ 19
3.1 BRIEF DESCRIPTION OF EXISTING SETUP, ADEKUS ............................................................................ 19
3.2 ADAPTATIONS FOR HK TURBINE TESTING ......................................................................................... 19
3.2.1 Alteration of flow conditions .......................................................................................................... 19
3.2.2 CFD simulation of the (open) channel flow .................................................................................... 27
4 BEM SIMULATION RESULTS ............................................................................................................. 33
4.1 SIMULATION METHOD ........................................................................................................................ 33
4.2 COMPUTATIONAL PROCEDURE, RESULTS AND DISCUSSION .............................................................. 33
5 DESIGN AND MANUFACTURING OF A LAB-SCALE HK-10 TURBINE ................................. 41
5.1 DESIGN AND MANUFACTURING OF PHYSICAL COMPONENTS ........................................................... 41
5.2 DATA ACQUISITION AND CONTROL SYSTEM ...................................................................................... 47
5.2.1 Measuring system .......................................................................................................................... 47
5.2.2 RPM Control system ...................................................................................................................... 49
6 EXPERIMENTAL STUDY ...................................................................................................................... 51
6.1 WIND TUNNEL EXPERIMENTS WITH HYDROKINETIC TURBINES ......................................................... 51
6.1.1 Brief description of the wind tunnel testing facility, UnB ............................................................. 51
6.1.2 Interpretation of wind tunnel results using dimensional analysis ................................................. 52
6.2 EXPERIMENTAL PROCEDURE, RESULTS AND DISCUSSION .................................................................. 55
7 CONCLUSIONS AND RECOMMENDATIONS .............................................................................. 59
REFERENCES .................................................................................................................................................... 63

xi
xii
1 INTRODUCTION
1.1 Background information
Humankind has been harnessing energy from flowing water to perform work for many
millennia. One of the first known applications of this energy source was grinding wheat
into flour by employing the mechanical energy extracted from water wheels in river
streams (Smith, 1975). The use of these hydraulic wheels and other rudimentary in-stream
devices has, however, been overwhelmed by the evolution of modern types of
hydropower turbines, which began in the 18th century. The driving factor for this evolution
was the need for large scale and more efficient energy production to accommodate the
rapidly increasing energy demand mainly due to the industrial revolution (Reynolds,
1983). This technological development was however characterized by the construction of
large hydroelectric dams and the submersion of extensive areas of land. Although these
dammed hydro power facilities have numerous advantages such as (Bagher, et al., 2015)
i) their ability to quickly adapt to changing energy demand, ii) their long economic lives
and low energy production costs, and iii) reduced CO2 emissions with respect to fossil fuel
based power production, there are some major drawbacks to be considered as well. For
example, the inherent need for flooding large areas of land to exploit this technology has
an environmental and social impact that is not to be underestimated; ecosystem damage,
possible loss of biologically rich and productive agricultural land, methane emissions due
to decaying plant material in an anaerobic environment, and relocation of people and
wildlife are some of the peculiarities that are, conservatively speaking, less favorable
(Bagher, et al., 2015; Chen, et al., 2015; Mol, et al., 2007). Note that said methane emissions
are particularly substantial in tropical regions, where deforestation was omitted prior to
the impoundment of the reservoir. In these cases, the greenhouse gas emissions may be
even higher than those from a conventional fossil fuel fired thermal generation plant, as
some studies have shown (DelSontro, et al., 2010; Giles, 2006). An excellent example of a
hydroelectric power facility that is characterized by practically all of the aforementioned
negative impacts is that of Suriname; roughly 160,000 hectares (about 1% of the country
area) of biologically valuable rainforest were flooded to give rise to the so-called
Brokopondo reservoir. Tragically, the installed capacity of this hydropower facility is only
189 MW and accordingly its power density in terms of submerged area per installed unit
of power is 847 ha/MW, which is, in this respect, one of the worst in the world (Ledec &
Quintero, 2003).

Notwithstanding the fact that the adoption of large power producing facilities, such as
dammed hydropower plants, have effectively targeted the need for increased energy
production, the uninterrupted availability of energy has become a serious concern from
an energy security point of view. An energy system where supply is heavily centralized is
very vulnerable to failures and is strategically inadvisable when considering the
possibility of adverse situations of techno-economic and geo-political nature (Sovacool,
2011; Goldthau & Sovacool, 2012; Alanne & Saari, 2006). Contrarily, the increased benefits
of decentralized/distributed energy production are becoming widely recognized; energy
resilience is positively affected, particularly when considering multiple (smaller)
generation sources in a diversified interconnected grid system. As a result, a significant
portion of research on energy generation systems is progressively being diverted to

16RET1012
1
relatively small sustainable alternatives. In this regard, several energy solutions are being
looked into, ranging from pico/micro hydropower plants, thermochemical conversion of
biomass (waste) products and small-scale heat and power generation with photovoltaic,
solar thermal and wind systems, just to mention some. Most of these technologies have
already a well-established theoretical framework and there is experience with their design,
modelling, optimization and experimental analysis. Another technology that is currently
being explored is that of small hydrokinetic power systems for river currents (Lago, et al.,
2010; Anyi & Kirke, 2010; Els & Brasil, 2015; Khan, et al., 2009). Although this technology
is not a new concept, its technical literature is still poorly populated.

As opposed to conventional hydropower systems, with hydrokinetic turbines there is no


need to build dams and penstocks to benefit from hydropower and consequently the
mentioned disadvantages of dammed hydropower are not present here. On the other
hand, it should be noted that the deployment of hydrokinetic technology is dependent on
the river capacity, which changes with seasons. Nevertheless, the use of hydrokinetic
turbines is potentially a viable solution for small remote communities situated alongside
river streams, which currently do not have (proper) access to electricity. Particularly in the
Amazon region and in the hinterlands of the Guyana’s there are numerous examples of
villages that might benefit from hydrokinetic technology. This is one of the main reasons
why Brazil is investing in research on river based hydrokinetic power systems and
currently seems to be the forerunner in this area (Els & Brasil, 2015; Els, et al., 2018).

Particularly in the last decade several hydrokinetic energy conversion applications have
been explored in Brazil. Most efforts that were made in this technological area focused on
the stand-alone applications for isolated communities in the Amazon region. Recently, this
focus area has been broadened by incorporating the hydrokinetic potential in downstream
channels of hydroelectric power plants (da Silva, et al., 2017), which led to the
establishment of the HydroK project (Brasil & Lavaquial, 2017); a research and
development project executed in partnership with two Brazilian universities, which was
sponsored by AES-Brazil to conduct applied research on propeller horizontal axis
hydrokinetic turbines. As part of said project, the HK-10 hydrokinetic turbine was recently
developed.

1.2 Problem definition & Objectives


The design of the rotor of the HK-10 turbine (see Figure 1.1) is the result of an extensive
hydrodynamic study, involving blade element momentum theory and several CFD
simulations, complemented by steady-state experiments in a wind tunnel at the University
of Brasilia (Brasil & Lavaquial, 2017; Brasil, et al., 2017). The use of a wind tunnel for the
experimental evaluation of the HK-10 turbine was made possible through the application
of a dimension analysis accounting for the difference in Reynolds number between the
flow in the wind tunnel and a water flow at design conditions. It should be noted that
wind tunnel testing was considered an option due to the unavailability of a suitable water
flume testing facility. Although there is a fairly good agreement between the obtained
experimental results and the numerical results from CFD simulations, it is necessary to
validate the (transposed) wind tunnel results (and, frankly, the CFD simulation results as
well) with results from actual (controlled) underwater experiments in order to better
assess the feasibility of the employed setup (and methods) at the University of Brasilia for

16RET1012
2
Figure 1.1 Photorealistic rendering of the HK-10 rotor design.

experimental analysis of hydrokinetic turbines. Anticipating this validation study, some


limitations can already be identified: i) due to the extremely large deviation existing
between the Reynolds numbers of the flow inside the wind tunnel and the actual flow
through hydrokinetic turbines, respectively, it is practically impossible to achieve
dynamic similitude, which is an important criterion when experimenting with lab-scale
models, ii) given the compressibility of air it is not possible to study transient behavior of
hydrokinetic turbines in a wind tunnel. The dynamic response of the turbine’s rotational
velocity on sudden load changes (or flow conditions), would be dissimilar for fluids with
different compressibility ratios, i.e., there is a time-scale difference due to the dampening
effect in air, which is not present in water due to its incompressibility. Note that
knowledge on the transient behavior of turbines is important, for example, when
designing and testing control systems, such as frequency and Maximum Power Point
Tracking (MPPT) controllers. In this respect, conveniently, the Anton de Kom University
of Suriname has an open-loop water flume testing facility in its hydrological laboratory
which has only sporadically been used in the past; it however does not meet the technical
requirements to test hydrokinetic turbines without some physical adaptations.

It is the objective of the present work to i) conceptualize the adaptations to the mentioned
water flume facility required for the purpose of experimentally studying the
hydromechanical behavior of horizontal axis hydrokinetic turbines, ii) to physically
develop a scaled model of the HK-10 hydrokinetic turbine suitable for underwater
experiments, equipped with sensors for torque and RPM measurements, and iii) to
perform an experimental analysis of the developed hydrokinetic turbine in the wind
tunnel in Brasilia as a baseline for comparison between available experimental
methodologies and for validation purposes. More specifically, said analysis compromises
of the experimental determination of the HK-10 turbine’s power curve, which relates the
power coefficient (𝐶𝑝 ) of the turbine to its tip speed ratio (𝜆).

1.3 Structure of report


This document is structured as follows. Firstly, in Section 2, a brief on hydrokinetic energy
conversion technologies is provided, followed by an extensive description of relevant

16RET1012
3
theory on the design and analysis of lift-based horizontal-axis flow turbines. Next, in
Section 0, technical details of the existing water flume testing facility at the Anton de Kom
University are presented and modifications are proposed in order to enable experimental
analysis of hydrokinetic turbines. By means of two independent computational methods
the flow conditions in the projected test section for hydrokinetic turbines are predicted. In
Section 4, the hydrodynamic performance of the HK-10 hydrokinetic turbine is studied
from a theoretical perspective by employing the blade element momentum (BEM)
method. In the following section the design and manufacturing of a lab-scale HK-10
turbine suitable for underwater experiments is presented, followed by a description of the
developed data acquisition and control system. In Section 6, the hydrodynamic
performance of the HK-10 turbine is experimentally investigated. Lastly, Section 7
recapitulates the work, and conclusions and recommendations for future research are
given herein.

16RET1012
4
2 RELEVANT THEORY
2.1 A brief on hydrokinetic turbines
Hydrokinetic energy conversion (HEC) technologies can be classified in three main
categories: i) axial-flow systems, ii) oscillating systems, and iii) cross-flow systems (Guney
& Kaygusuz, 2010). Of course, all these systems have their advantages, disadvantages and
different areas of application. It is however out of the main interest of the present work to
elaborate on oscillating and cross-flow technologies; the information presented in this
section will be focusing on lift-based, axial flow turbines; the HK-10 hydrokinetic turbine
introduced earlier falls within this category. These turbines are also known as zero-head or
in-stream turbines, because of their ability to produce power without the need of a head
differential, i.e., instead of potential energy conversion, such as in dammed-hydropower
plants, it is the kinetic energy that is converted from, for example, ocean or river streams
which drives the hydrokinetic turbine.

The working principle of lift-based axial-flow turbines is basically the same as that of
typical wind turbines and hence much of the developed theory on these energy conversion
machines has been derived from existing wind turbine theory. The rotor blades of a
hydrokinetic turbine are composed of airfoil-shaped 2-D cross sections, such as shown in
Figure 2.1, which is the one used in the design of the HK-10 turbine (NACA 4415 airfoil).
0.18

0.09

0.00 chord line

-0.09
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0

Figure 2.1 – The NACA 4415 airfoil (non-dimensional), used in the design of the HK-10 turbine. The geometry
of the airfoil is depicted from its leading edge (0, 0) up to its trailing edge (1, 0). Coordinates have been obtained
from the UIUC airfoil database (Anon., n.d.).

Given specific operating conditions, the optimal rotor geometry of a lift-based horizontal
axis turbine can be computed using Blade Element Momentum (BEM) theory (Hansen,
2008). Note that the performance of these turbines is generally expressed in terms of their
power coefficient, 𝐶𝑃 , which is a non-dimensional parameter defined as the ratio of the
(mechanical) power extracted from the turbine, 𝑃, to the available kinetic power contained
in the flow through a cross-section equal to the swept area, 𝐴, by the rotor (Hansen, 2008),
i.e.,
𝑃
𝐶𝑃 ≡ (2.1)
1 3
2 𝜌𝑢∞ 𝐴

16RET1012
5
where 𝜌 is the density of the fluid and 𝑢∞ is the flow speed far upstream of the rotor in a
direction which is parallel to its shaft.

It is important to explicitly state at this point that the power coefficient does not have one
fixed value for a specific turbine; it is rather dependent on the operating conditions, which
is clearly evident from equation (2.1). These operating conditions affect the rotational
speed of the rotor and therefore, conveniently, the power coefficient is usually related to
a so-called Tip Speed Ratio, 𝜆, defined as follows (Hansen, 2008),
𝜔𝑅
𝜆≡ (2.2)
𝑢∞
where 𝜔 and 𝑅 denote the rotational velocity of the rotor and its radius, respectively.

In Table 2.1 some design characteristics and the (projected) nominal operating conditions
of the HK-10 turbine are presented.

Table 2.1 – Technical specifications of the HK-10 turbine and its nominal operating conditions (Brasil &
Lavaquial, 2017).

Diameter, D [m] 2.1


Hydrofoil NACA 4415
Number of blades, 𝑁𝐵 4
Rotational speed, n [RPM] 35
Water flow speed, 𝑢∞ [m/s] 2.5
Tip Speed Ratio, 𝜆 [-] 1.61
Power coefficient, 𝐶𝑃 [-] 0.4

In the next sections the dynamics of (hydro)kinetic turbines are discussed in detail,
introducing the actuator disc model and Blade Element Momentum (BEM) theory as
powerful tools to analyze said turbines from a theoretical point of view.

2.2 Rankine-Froude actuator disc model


The Rankine-Froude actuator disc model is the simplest model to describe the dynamics
of horizontal axis (hydro)kinetic turbines1 (Hansen, 2008). In this model the rotor of the
hydrokinetic turbine is modelled as permeable disc, through which there is flow in only
one direction, i.e., there is no rotational velocity component in its wake. Moreover, it is
assumed that the disc operates frictionless and consequently it is considered ideal. In
principle, in this model the rotor disc acts as a device which slows down the water speed
from far upstream of the rotor, 𝑢∞ , to a velocity 𝑢𝑑 , at the rotor plane. This decrease in
velocity is induced by an increase in pressure close upstream of rotor, from 𝑝0 to 𝑝𝑑+ . Note
that here 𝑝0 equals the atmospheric pressure added by the hydrostatic pressure far
upstream of the rotor. At the rotor disc itself the pressure experiences a discontinuous

1Both the Rankine-Froude model and the BEM model discussed in the next section are typically
employed for the analysis of wind turbines, where the blade length is typically much larger than
the chord width, often referred to as slender blade turbines. There are, however, numerous
examples of applications with wide blade rotors, such as marine propellers and hydrokinetic
turbines, where excellent performance (of the mentioned models) is shown (Masters, et al., 2011;
Benini, 2004; Kolekar, et al., 2011). 6

16RET1012
drop, from 𝑝𝑑+ to 𝑝𝑑− , as opposed to the flow velocity which continuously (provided that
density of the fluid is constant) decreases further to 𝑢𝑤 in the wake. The pressure, on the
other hand, gradually increases back in the wake until it returns to 𝑝0 . This behavior of
velocity and pressure is depicted graphically in Figure 2.2.

The (discontinuous) pressure drop from 𝑝𝑑+ to 𝑝𝑑− is responsible for a force in the
streamwise direction, acting on the actuator disc, also known as the thrust, 𝑇,

𝑇 = ∆𝑝 ∙ 𝐴 (2.3)
Assuming that the flow through the turbine rotor is stationary, incompressible and
frictionless, and there are no external forces acting on the fluid either upstream or down-

(a)

(b)

(c)
Figure 2.2 (a) The rotor of the hydrokinetic turbine is modelled as an actuator disc, enclosed by a control
volume which is delineated by the outer streamlines shown. (b) Velocity profile from far upstream to far
downstream of the actuator disc. (c) Pressure profile from far upstream to far downstream of the actuator disc.
There is a discontinuous pressure drop across the actuator disc. Adapted from Ref. (Hansen, 2008).

16RET1012
7
stream of the rotor, the Bernoulli equation (equation (2.4)) can be employed to describe
the hydrodynamic behavior of the fluid flow (White, 2015),

𝑝1 1 2 𝑝2 1
+ 𝑉1 + 𝑔𝑧1 = + 𝑉22 + 𝑔𝑧2 = constant (2.4)
𝜌 2 𝜌 2
Where 𝜌 denotes the mass density of the fluid, 𝑉 is the velocity of the fluid, 𝑧 is the
elevation with respect to a reference height, and subscripts (1) and (2) respectively indicate
inlet and outlet properties given a defined control volume (CV).

Applying the Bernoulli equation respectively to a CV ranging from far upstream to just in
front of the rotor and another CV ranging from just behind the rotor to far downstream,
yields,
1 2 1
𝑝0 + 𝜌𝑢∞ = 𝑝𝑑+ + 𝜌𝑢𝑑2 (2.5)
2 2
and
1 1 2
𝑝𝑑− + 𝜌𝑢𝑑2 = 𝑝0 + 𝜌𝑢𝑤 (2.6)
2 2
By combining equations (2.5) and (2.6), the following relation can be obtained:
1
𝑝𝑑+ − 𝑝𝑑− = ∆𝑝 = 𝜌(𝑢∞
2 2)
− 𝑢𝑤 (2.7)
2
Application of the axial momentum equation to a cylindrical control volume such as
depicted in Figure 2.3, taking into account all of the previously mentioned assumptions
leads to,
2 2 (𝐴 2
𝜌𝑢𝑤 𝐴𝑤 + 𝜌𝑢∞ CV − 𝐴𝑤 ) + 𝑚̇side 𝑢∞ − 𝜌𝑢∞ 𝐴CV = −𝑇 (2.8)
where 𝐴CV equals the area of the circular cross section of the control volume, 𝐴𝑤 is the
cross section of the streamtube in the wake and 𝑚̇𝑠𝑖𝑑𝑒 denotes the mass flow through the
lateral boundary of the control volume. Note that since the pressure has the same
magnitude on both the inlet and the outlet of the CV, and there are no axial pressure forces
components acting on the lateral boundary, only the thrust force, 𝑇, is involved here.

Figure 2.3 Cylindrical control volume around a hydrokinetic turbine. Adapted from Ref. (Hansen, 2008).

16RET1012
8
The mass flow through the lateral boundary of the control volume, 𝑚̇side , can be
determined from the mass balance,

𝜌𝐴𝑤 𝑢𝑤 + 𝜌(𝐴CV − 𝐴𝑤 )𝑢∞ + 𝑚̇side = 𝜌𝐴CV 𝑢∞ (2.9)


which results in,
𝑚̇side = 𝜌𝐴𝑤 (𝑢∞ − 𝑢𝑤 ) (2.10)
Moreover, application of the mass balance also gives a relationship between 𝐴 and 𝐴𝑤 ,

𝑚̇ = 𝜌𝑢𝑑 𝐴 = 𝜌𝑢𝑤 𝐴𝑤 (2.11)


Note that here 𝑚̇ is the mass flow through the streamtube enclosed by the solid lines in
Figure 2.3.

Combination of equations (2.8), (2.10) and (2.11) leads to:

𝑇 = 𝜌𝑢𝑑 𝐴(𝑢∞ − 𝑢𝑤 ) = 𝑚̇(𝑢∞ − 𝑢𝑤 ) (2.12)


Furthermore, combination of equations (2.3), (2.7) and (2.12) results in a very intriguing
finding:
1
𝑢𝑑 = (𝑢∞ + 𝑢𝑤 ) (2.13)
2
Equation (2.13) basically states that the axial flow velocity at the rotor plane is the mean
of flow velocity far upstream and the flow velocity in the wake.

In order to derive an expression from which the turbine’s shaft power, 𝑃 , can be
computed, the energy conservation law is applied to a control volume which coincides
with the stream tube enclosed by the solid lines in Figure 2.3, yielding
1 2 𝑝0 1 2 𝑝0
𝑃 = 𝑚̇ ( 𝑢∞ + − 𝑢𝑤 − ) (2.14)
2 𝜌 2 𝜌
Since 𝑚̇ = 𝜌𝑢𝑑 𝐴,
1 2 2)
𝑃 = 𝜌𝑢𝑑 𝐴(𝑢∞ − 𝑢𝑤 (2.15)
2
In the subsequent analysis the axial induction factor, 𝑎 , will be frequently used; this
quantity is defined as follows:
𝑢∞ − 𝑢𝑑
𝑎≡ (2.16)
𝑢∞
By combining equation (2.13) and (2.16) the flow velocity in the wake of the turbine can
be expressed as
𝑢𝑤 = (1 − 2𝑎)𝑢∞ (2.17)
and consequently
3
𝑃 = 2𝜌𝑢∞ 𝑎(1 − 𝑎)2 𝐴 (2.18)
Substitution of equation (2.18) into equation (2.1) gives an expression for the (non-
dimensional) power coefficient in terms of the axial induction factor:

𝐶𝑃 = 4𝑎(1 − 𝑎)2 (2.19)


This relationship is illustrated in Figure 2.4.

16RET1012
9
0.7

0.6 Betz limit = 16⁄27 ≈ 0.593

0.5

0.4
CP [-]

0.3

a = 1/3
0.2

0.1

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
a [-]

Figure 2.4 Relationship between the power coefficient, 𝐶𝑃 , and the axial induction factor, 𝑎. It can be seen from
the graph that 𝐶𝑃 reaches a maximum value of approximately 0.593, corresponding to the Betz limit.

𝐶𝑃 has a theoretical limit, generally known as the Betz limit, which is determined by
differentiating equation (2.19) with respect to a:

𝑑𝐶𝑃
= 4(1 − 𝑎)(1 − 3𝑎) (2.20)
𝑑𝑎
Solving 𝑑𝐶𝑃 ⁄𝑑𝑎 = 0, gives 𝐶𝑃, max = 16⁄27 ≈ 0.593, corresponding to 𝑎 = 1⁄3 (see also
Figure 2.4)

2.3 Effects of rotation


In the previous paragraph the effect of wake rotation was not considered as the turbine
rotor was considered ideal. However, in reality, the rotation of the rotor induces angular
velocity components in the wake of the turbine in a direction opposite to that of the turbine
blades (Hansen, 2008). These angular velocity components are usually quantified in terms
of an angular induction factor, 𝑎′, which is defined as follows:
𝑢𝜃,𝑤
𝑎′ (𝑟) ≡ (2.21)
2𝜔𝑟
In equation (2.21) 𝑢𝜃,𝑤 denotes the tangential velocity component in the wake, 𝜔 is the
rotational speed of the rotor, whereas 𝑟 represents the perpendicular distance between
rotor’s axis of rotation and an arbitrary point in the wake just behind actuator disc. Note
that the angular induction factor is inversely proportional to 𝑟 and therefore wake rotation
is more pronounced close the rotor’s hub.

16RET1012
10
The tangential velocity of the flow at the actuator disc, 𝑢𝜃 , is defined as the sum of the
tangential velocity of the rotor and the average between the tangential velocity
components of the flow immediately before and after the actuator disc (resp. zero and
𝑢𝜃,𝑤 ), thus

𝑢𝜃 = 𝜔𝑟(1 + 𝑎′) (2.22)


The axial induction factor and the angular induction factor are not independent for local
angles of attack, 𝛼, below stall. In fact, the induced velocity, 𝑤 ⃗⃗ , must be in the same
direction as the lift force on the blade and thus perpendicular to the local velocity. From
velocity triangles depicted in Figure 2.5 it can be easily seen that

𝑎′𝜔𝑟 (1 − 𝑎)𝑢∞ (1 − 𝑎)
tan 𝜙 = = = (2.23)
𝑎𝑢∞ (1 + 𝑎′)𝜔𝑟 (1 + 𝑎′)𝜆𝑟
where 𝜆𝑟 = 𝜔𝑟/𝑢∞ denotes the ratio between the local rotational speed of the rotor and
the flow speed.

From equation (2.23) a relation between 𝑎 and 𝑎′ can be derived, namely,

𝜆2𝑟 𝑎′ (1 + 𝑎′ ) = 𝑎(1 − 𝑎) (2.24)


which can be solved for 𝑎′ resulting in

1 1
𝑎′ = √1 + 4𝜆−2
𝑟 𝑎(1 − 𝑎) − (2.25)
2 2
The induced angular velocity components affect the performance of the turbine as was
previously pointed out. In order to derive an expression for the power coefficient, 𝐶𝑃 , of
the turbine which takes into account said influence, the actuator disc is divided into
infinitesimally small rotating annuli (Hansen, 2008). The torque, 𝑇, exerted by the flow on

Figure 2.5 Velocity triangles on a hydrofoil. By definition, the lift force, L, is perpendicular to the direction of
the incoming flow. Adapted from Ref. (Hansen, 2008).

16RET1012
11
an annulus of thickness 𝑑𝑟 and located at a distance 𝑟 from the turbines axis of rotation is
given by
𝑑𝑇 = 4𝜋𝜌𝜔𝑢∞ (1 − 𝑎)𝑎′𝑟 3 𝑑𝑟 (2.26)
Consequently, the shaft power contribution, 𝑑𝑃, of the flow through the annulus is

𝑑𝑃 = 𝜔 ∙ 𝑑𝑇 = 4𝜋𝜌𝜔2 𝑢∞ (1 − 𝑎)𝑎′𝑟 3 𝑑𝑟 (2.27)


The total shaft power, 𝑃, is computed by integrating 𝑑𝑃 from 𝑟 = 0 to 𝑟 = 𝑅, where 𝑅
equals the radius of the rotor:
𝑅
𝑃 = 4𝜋𝜌𝜔2 𝑢∞ ∫ (1 − 𝑎)𝑎′𝑟 3 𝑑𝑟 (2.28)
0

Equation (2.28) can be written in dimensionless form as

8 𝜆
𝐶𝑃 = ∫ (1 − 𝑎)𝑎′𝜆3𝑟 𝑑𝜆𝑟 (2.29)
𝜆2 0
Note that the integral is generally computed numerically by dividing the rotor blades into
𝑁 sections (ranging from the rotor hub to the blade tip) with a thickness of Δ𝑟.

2.4 Optimal flow conditions


Given nominal operating conditions for the hydrokinetic turbine (𝑢∞ , 𝜔, 𝑅), it is possible
to compute optimal velocity triangles along the turbine blades in order to maximize its
power coefficient. From equations (2.28) and (2.29) it clear that to optimize the turbine’s
shaft power it is necessary to maximize the following expression (Hansen, 2008):

𝑓(𝑎, 𝑎′) = 𝑎′(1 − 𝑎) (2.30)


Since 𝑎′ is a function 𝑎 (recall equation (2.25)), the optimization problem may be
formulated as 𝑑𝑓⁄𝑑𝑎 = 0. Application of the chain rule yields:

𝑑𝑓 𝑑𝑎′
= (1 − 𝑎) − 𝑎′ = 0 (2.31)
𝑑𝑎 𝑑𝑎
On the other hand, differentiating both sides of equation (2.24) with respect to 𝑎 gives:

𝑑𝑎′ 2
(1 + 2𝑎′) 𝜆 = 1 − 2𝑎 (2.32)
𝑑𝑎 𝑟
The optimal relationship between 𝑎 and 𝑎′ is found by combining equations (2.24), (2.31)
and (2.32):
1 − 3𝑎
𝑎′ = (2.33)
4𝑎 − 1
By equating the right hand sides of equations (2.25) and (2.33) to each other it is possible
to compute the optimal axial induction factors (and therefore also the optimal angular
induction factors) numerically:

1 − 3𝑎 −1 + √1 + 4𝜆−2
𝑟 𝑎(1 − 𝑎)
= (2.34)
4𝑎 − 1 2

16RET1012
12
0.7

0.6
Betz limit = 16/27 ≈ 0.593

0.5

0.4
CP, max [-]

0.3

0.2

0.1

0.0
0 1 2 3 4 5 6 7
TSRnom [-]

Figure 2.6 Maximum 𝐶𝑃 values for different tip-speed ratios, respectively, when considering the influence of
wake rotation. The power coefficient asymptotically approaches the Betz limit as tip-speed-ratios increase to
infinitely large values.

Anticipating the numerical results from the BEM computations in the next section, Figure
2.6 is presented here showing maximum 𝐶𝑃 values for different nominal tip-speed-ratios
(without considering blade tip and hub losses due to complex flow phenomena) for the
HK-10 rotor. As is evident from Figure 2.6, the power coefficient asymptotically
approaches the Betz limit as tip-speed-ratios increase to infinitely large values.
Consequently, from an energy efficiency point of view it is desirable for the turbine to
operate at high rotational speeds in order to minimize losses due to a rotating wake.

2.5 Blade Element Momentum Theory


The Blade Element Momentum (BEM) method, developed by Froude in 1878 and refined
by Glauert in 1926, relates momentum theory to local events taking place at the rotor
blades (Hansen, 2008). Application of this method involves discretization of the (variable
diameter) stream tube, which was previously introduced, into annular elements of
thickness 𝑑𝑟, whereby it is assumed that: i) there is no radial dependency between annular
elements, ii) the flow experiences a force from the rotor blades which is constant for every
annular element, given a fixed operating condition.

With the application of the integral momentum equation to an annular control volume
within the stream tube it is possible to determine the thrust from the actuator disc (on the
control volume considered):

𝑑𝑇 = (𝑢∞ − 𝑢𝑤 )𝑑𝑚̇ = 2𝜋𝑟𝜌𝑢𝑑 (𝑢∞ − 𝑢𝑤 )𝑑𝑟 (2.35)

16RET1012
13
On the other hand, the torque, 𝑀, on the annular control volume is found by employing
the angular momentum equation:

𝑑𝑀 = 𝑟𝑢𝜃,𝑤 𝑑𝑚̇ = 2𝜋𝑟 2 𝜌𝑢𝑑 𝑢𝜃,𝑤 𝑑𝑟 (2.36)


Using the definitions for the axial and angular induction factors in equations (2.16) and
(2.21), respectively, together with equation (2.17), the thrust and torque on the annular
control volume can be reformulated as:
2
𝑑𝑇 = 4𝜋𝑟𝜌𝑢∞ 𝑎(1 − 𝑎)𝑑𝑟 (2.37)
and
𝑑𝑀 = 4𝜋𝑟 3 𝜌𝑢∞ 𝜔(1 − 𝑎)𝑎′𝑑𝑟 (2.38)
Alternatively, by considering local events (e.g., acting hydrodynamic forces and flow
velocities) taking place at a section of the rotor blade contained within the annular control
volume, the thrust and torque can also be computed, which is the focus of the following
discussion.

Recall from basic aerodynamics theory that lift force, 𝐿, on an airfoil is defined such that
it is always perpendicular to the fluid velocity as seen by the foil, whereas the drag force,
𝐷, is parallel to this velocity (see also Figure 2.7). The lift and drag force per length can be
expressed as:
1 2
𝐿 = 𝜌𝑢∞ 𝑐𝐶𝐿 (2.39)
2
and
1 2
𝐿 = 𝜌𝑢∞ 𝑐𝐶𝐷 (2.40)
2
where 𝐶𝐿 and 𝐶𝐷 are respectively the lift and drag coefficient, and 𝑐 denotes the chord
length. Note that said coefficients are strongly dependent on the shape of the foil, the
Reynolds number of the flow and the angle of attack, 𝛼, as is defined in Figure 2.8.

Figure 2.7 Local forces acting on a blade section. Adapted from Ref. (Hansen, 2008).

16RET1012
14
Figure 2.8 Velocities at the rotor plane. Adapted from Ref. (Hansen, 2008).

Projection of the lift and drag forces in the normal and tangential direction with respect to
the rotor plane gives (see Figure 2.7):

𝑃𝑁 = 𝐿 cos 𝜙 + 𝐷 sin 𝜙 (2.41)


and
𝑃𝑇 = 𝐿 sin 𝜙 − 𝐷 cos 𝜙 (2.42)

Following from the fact that both 𝑃𝑁 and 𝑃𝑇 are forces per unit length of the rotor blades,
the thrust and torque from the annular control volume previously considered equals

𝑑𝑇 = 𝑃𝑇 𝑁𝐵 𝑑𝑟 (2.43)
and
𝑑𝑀 = 𝑟𝑃𝑇 𝑁𝐵 𝑑𝑟 (2.44)
Note that here 𝑁𝐵 represents the number of blades.
2
Normalization of equations (2.41) and (2.42) with respect to (1⁄2 𝜌𝑢𝑟𝑒𝑙 𝑐) respectively
yields

𝐶𝑁 = 𝐶𝐿 cos 𝜙 + 𝐶𝐷 sin 𝜙 (2.45)


and
𝐶𝑇 = 𝐶𝐿 sin 𝜙 − 𝐶𝐷 cos 𝜙 (2.46)
where
𝑃𝑁
𝐶𝑁 = 2 (2.47)
1⁄2 𝜌𝑢𝑟𝑒𝑙 𝑐
and
𝑃𝑇
𝐶𝑇 = 2 (2.48)
1⁄2 𝜌𝑢𝑟𝑒𝑙 𝑐
Furthermore, it can be derived from Figure 2.8 that

𝑢𝑟𝑒𝑙 sin 𝜙 = 𝑢∞ (1 − 𝑎) (2.49)


and
𝑢𝑟𝑒𝑙 cos 𝜙 = 𝜔𝑟(1 + 𝑎′) (2.50)
At this point equations (2.43) and (2.44) can be rewritten by combining them with
equations (2.47) and (2.49), and equations (2.48), (2.49) and (2.50), respectively, yielding
2 (1
1 𝑢∞ − 𝑎)2
𝑑𝑇 = 𝜌𝑁𝐵 𝑐𝐶𝑁 𝑑𝑟 (2.51)
2 sin2 𝜙

16RET1012
15
and
1 𝑢∞ (1 − 𝑎)𝜔𝑟(1 + 𝑎′)
𝑑𝑀 = 𝜌𝑁𝐵 𝑐𝐶𝑇 𝑟𝑑𝑟 (2.52)
2 sin 𝜙 cos 𝜙
Expressions for the axial and angular induction factors in terms of local quantities can be
obtained by equalizing equations (2.37) and (2.51), and equations (2.38) and (2.52):
−1
8𝜋𝑟 sin2 𝜙
𝑎=( + 1) (2.53)
𝑐𝑁𝐵 𝐶𝑁
and
−1
8𝜋𝑟 sin 𝜙 cos 𝜙
𝑎′ = ( − 1) (2.54)
𝑐𝑁𝐵 𝐶𝑇
In principle, all the necessary equations to perform basic BEM computations to yield
results regarding normal and tangential load distributions have been presented.
Moreover, by integrating the normal and tangential loads from all blade elements
considered it is possible to determine the turbine’s thrust and the shaft power, which is
probably the most important outcome of such analysis. However, in order to obtain
improved results, two corrections must be applied, namely, i) Prandtl’s tip/hub loss
correction, and ii) Glauert’s correction for high axial induction values.

The classical BEM method (Rankine, 1865; Froude, 1878; Glauert, 1983; Hansen, 2008)
presented until now assumed a rotor with an infinite number of blades. The vortex in the
wake of the turbine is, however, different for a rotor consisting of a finite number of
blades, which affects hydrodynamic forces. To account for this, Prandtl has introduced a
correction factor, 𝐹, in equations (2.37) and (2.38), namely
2
𝑑𝑇 = 4𝜋𝑟𝜌𝑢∞ 𝑎(1 − 𝑎)𝐹𝑑𝑟 (2.55)
and
𝑑𝑀 = 4𝜋𝑟 3 𝜌𝑢∞ 𝜔(1 − 𝑎)𝑎′𝐹𝑑𝑟 (2.56)
Said correction factor is often referred to as Prandtl’s tip/hub loss factor and is formulated
as follows (Hansen, 2008; Shen, et al., 2005)

𝐹 = 𝐹𝑡𝑖𝑝 𝐹ℎ𝑢𝑏 (2.57)


where
2 𝑁 (𝑅−𝑟)
− 𝐵
𝐹𝑡𝑖𝑝 = cos−1 [𝑒 2 𝑟 sin 𝜙 ] (2.58)
𝜋
and
2 𝑁 (𝑟−𝑅ℎ )
−1 − 𝐵
𝐹ℎ𝑢𝑏 = cos [𝑒 2 𝑟 sin 𝜙 ] (2.59)
𝜋

In equation (2.59) 𝑅ℎ refers to the hub radius.

Employing equations (2.55) and (2.56) instead of equations (2.37) and (2.38) in order to
derive expressions for the axial and angular inductions factors gives
−1
8𝜋𝑟𝐹 sin2 𝜙
𝑎=( + 1) (2.60)
𝑐𝑁𝐵 𝐶𝑁

16RET1012
16
and
−1
8𝜋𝑟𝐹 sin 𝜙 cos 𝜙
𝑎′ = ( − 1) (2.61)
𝑐𝑁𝐵 𝐶𝑇
Note that Prandtl’s tip/hub loss factor also affects equations (2.51) and (2.52).

It has been shown for wind turbines that momentum theory breaks down when axial
induction factors become larger than approximately 0.45 (= 𝑎𝑐𝑟𝑖𝑡 ), i.e., equation (2.60) has
limited validity (Hansen, 2008). In principle, whenever high axial induction factors are
observed during calculations, empirical relations based on experiments are used instead
of equation (2.60). Herein, an empirical formulation proposed by Buhl (Chapman, et al.,
2013) is employed for 𝑎 > 𝑎𝑐𝑟𝑖𝑡 :

1
𝑎 = [2 + (𝐾(1 − 2𝑎𝑐𝑟𝑖𝑡 ) + 2)2 − √𝐾(1 − 2𝑎𝑐𝑟𝑖𝑡 ) + 4(𝐾𝑎𝑐𝑟𝑖𝑡
2
− 1)] (2.62)
2
where
8𝜋𝑟𝐹 sin2 𝜙
𝐾= (2.63)
𝑐𝑁𝐵 𝐶𝑁

This section is concluded with an expression to calculate the chord length of a blade
section, based on optimal flow conditions, and taking into account Prandtl losses.
Combination of equations (2.37) and (2.51) gives:

8𝜋𝑟𝑎𝐹 sin2 𝜙
𝑐(𝑟) = (2.64)
𝑁𝐵 (1 − 𝑎)𝐶𝑁

16RET1012
17
16RET1012
18
3 WATER FLUME TESTING FACILITY
In the present section a comprehensive description of the suggested adaptations to the
existing water flume testing facility in the hydrological laboratory at the AdeKUS is
provided. These adaptations are considered necessary to conduct controlled underwater
experiments with hydrokinetic turbines. Results from both application of the energy
conservation method, as well as CFD computations are presented herein to elucidate on
the flow conditions in the projected test section for hydrokinetic turbines. These
conditions, in particular the water level and velocity distribution, have been important
inputs with respect to scaling of the HK-10 turbine for experimental study.

3.1 Brief description of existing setup, AdeKUS


In the hydrological laboratory of the AdeKUS there is a water flume testing facility, which
was built in 2006 with the purpose to conduct hydraulic and hydrologic experiments.
Furthermore, the facility enables experiments with micro and pico hydropower (reaction)
turbines.

The experimental setup consist of large water reservoir with a capacity of 120 m3 at the
bottom, from which water is pumped to a smaller reservoir of 60 m3 at the top, using two
centrifugal pumps with a pumping capacity of 400 l⁄s , each (2x 121 kW). The top
reservoir is located at a height of approximately 10 m with respect to the bottom reservoir.
Water leaves the top reservoir through a discharge pipe equipped with a magnetic flow
sensor and throttle mechanisms to control the discharge rate and to eventually divert the
water flow to either the water flume (for hydraulic/hydrologic experiments) or the inlet of
an installed reaction turbine. Moreover, there is an overflow pipe which prevents the top
reservoir from overflowing from its top edges in some conditions.

The water flume has a length of 22 m, a width of 1 m, a depth of 1.25 m and features
adjustable inclination of its side walls. Water is fed into the flume almost perpendicularly
by means of two pipes (𝐷 = 0.3 m) as is shown in Figure 3.1. From experiments that have
been conducted in the past, it was concluded that the design of the water supply induces
an extremely turbulent flow in the flume. With the use of an improvised honeycomb
straightener built from PVC pipes it was attempted to create a laminar condition, which
was only moderately successful. Note that water from the flume flows back to the bottom
reservoir from which it is pumped up again, and consequently the experimental setup
may be considered closed-loop system.

3.2 Adaptations for HK turbine testing


3.2.1 Alteration of flow conditions
Since hydrokinetic turbines are machines that extract kinetic energy from water flows, the
employed experimental setup should be designed such that kinetic energy losses are
minimized. Moreover, testing the turbine at higher velocities is advantageous with respect
to homogenization of the fluid velocity vector field, i.e., boundary effects are less
perceptible. Taking this into account, it is best to not perform any throttling in the
discharge from the top reservoir into the flume, as long as a stationary water level inside
the flume can be achieved. Moreover, it is proposed to modify the water supply piping
system in such a way, that water is fed into the flume tangentially with respect to its bed,

16RET1012
19
Figure 3.1 Depiction of the water flume testing facility at the Anton the Kom University of Suriname. As can
be seen, water is fed almost perpendicularly into the flume.

to reduce energy losses and unfavorable velocity components in the flume.

Assuming that i) the water discharge system allows for a flow magnitude equal to that
provided by the centrifugal pumps, ii) kinetic energy losses during supply of water into
the flume are negligibly small, and iii) wall friction can be neglected, it can be easily shown
that the water inside the flume should reach a level of about 14 cm, whereas the (average)
flow velocity in the direction parallel to the flume’s bed would be 5.6 m⁄s. It can be argued
that a water level of only 14 cm is not practical for testing hydrokinetic turbines; to avoid
blockage effects and boundary layer influence the scale model of the turbine would need
to be very small. Although manufacturing such a small model is technically feasible, it is
not to be recommended; while turbulence carries vortices from the Smagorinsky scale up
to the length scale of the flow geometry, a (possible) wavy flow may result in a non-
homogenous velocity field and unbalanced pressure distribution on the rotor, namely
when the transverse motion of the fluid is significantly greater than the rotor size.

In order to alter the water level in the flume two options have been considered: i) by means
of a barrier or weir plate (e.g., V-notch or rectangular type) across the width of the flume,
and ii) by means of a channel transition or so-called hydraulic jump, which is only possible
if the upstream flow is supercritical. To achieve supercritical conditions, the Froude
number should be larger than unity, Fr > 1 (Chaudhry, 2007). In fluid mechanics, the
Froude number is a dimensionless quantity defined as the ratio of the flow inertia to the
external field or gravity field in this case, i.e. (Chaudhry, 2007; White, 2015),
𝑢1
Fr = (3.1)
√𝑔𝑦1

16RET1012
20
where 𝑢0 refers to the upstream (before the transition) flow velocity and 𝑦1 represents the
upstream water level. With 𝑢1 = 5.6 m⁄s , 𝑦1 = 0.14 m and 𝑔 = 9.81 m⁄s 2 , the Froude
number is calculated to be Fr = 4.78. Note that this number only corresponds to a scenario
where no throttling is applied in the discharge pipes coming from the top reservoir.

The option of employing a weir plate to increase the water level in the flume was
dismissed as it was argued that large velocity gradients in the vertical direction might be
present near the end of the flume, which is where testing of the hydrokinetic turbine
should be ideally performed when considering the entrance length of a fully developed
laminar flow. Instead, the option of using a channel transition for said purpose was
scrutinized. In principle, two types of transitions are possible (Chaudhry, 2007): i) constant
flume width transitions (e.g., a step-up or step-down in the flume bed) and ii) variable
flume width transitions (e.g., a contraction or expansion of the flume). If a rise in the water
level is desired, a step-up or contraction in the flume would be necessary. For the present
case it was decided to design a gradual contraction in the flume’s width, as this should
promote homogenization of the velocity profile across the test section for the turbine,
which is advantageous.

The analysis for channel transitions is generally done by enforcing the law of energy
conservation. For a rectangular (cross-section) channel with a hydrostatic pressure
distribution and uniform velocity distribution, the specific energy at some point in the
channel is expressed as (Chaudhry, 2007):

𝑞2
𝐸 =𝑦+ (3.2)
2𝑔𝑦 2
where 𝑦 is the flow depth and 𝑞 is defined as the discharge, 𝑄, per unit width. Assuming
a frictionless flow, the specific energy remains constant throughout the length of the
channel, whether or not channel transitions are present. Several interesting properties of
channel transitions can be derived from equation (3.2) when performing basic calculus
operations. Some of these properties are (the derivation is omitted here):

3 𝑞2 2
𝑦𝑐 = √ = 𝐸 (3.3)
𝑔 3
𝑢𝑐
Fr = =1 (3.4)
√𝑔𝑦𝑐
2
8
𝑞𝑚𝑎𝑥 = 𝑔𝐸 3 (3.5)
27
In equation (3.4), 𝑢 refers to the flow velocity. Note that mathematical symbols with
subscript “c” indicate critical state properties. The critical state is defined as the state at
which the specific energy, 𝐸, is minimum for a given unit discharge, 𝑞.

With equations (3.2 - 3.5) the variation of unit discharge with flow depth for a range of
different values for the specific energy and channel width, 𝐵, can be computed. Processing
of such results into a 𝑦 − 𝑞 diagram provides insightful information regarding possible
channel transitions. For the project at hand, aforesaid diagram has been computed, see
Figure 3.2. Note that the 𝐸 = 𝐸𝑚𝑎𝑥 line corresponds to the flow state where no throttling
is performed in the discharge pipes from the top reservoir of the test facility, whereas all

16RET1012
21
1.8
B = 0.5m
1.6
B = 0.6m B = 0.4m
B = 0.3m
1.4
B = 0.7m

1.2 B = 0.2m

B = 0.8m
1.0
y [m]

B = 0.9m
0.8

0.6 E = Emax

0.4

0.2

0.0
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5
q [m3/m]

Figure 3.2 Variation of the flow depth with the unit discharge for different values of the specific energy and
different channel widths. Alteration of the flow depth is only possible for transitions below the dash-dotted
line.

other solid lines are iso-𝐸 lines where 𝐸 < 𝐸𝑚𝑎𝑥 , with throttling performed. The dashed-
dotted line represents the critical state, for which 𝑑𝐸 ⁄𝑑𝑦 = 0. As can be easily seen from
Figure 3.2, a transition from an arbitrary width 𝐵1 to a width 𝐵2 along a iso-𝐸 line, where
𝐵1 > 𝐵2 , leads to a rise in flow depth, only for flow states below the Fr = 1 line, i.e.,
alteration of the flow depth is only possible for a supercritical upstream flow, as was
indicated before. Subscripts (1) and (2) denote the pre- and post-transition states,
respectively.

Taking into account i) the range of velocities and flow depths at which the hydrokinetic
turbine can be tested, ii) minimization of the hydraulic radius for flow efficiency, and iii)
the introduction of velocity components in the direction perpendicular to rotor’s axis of
rotation due to the channel transition, it was concluded that a width transition to 0.5 m
provides a relatively favorable test condition, with respect to the other widths considered
in Figure 3.2. In order to gain a better understanding of the behavior of the flow velocity
and flow depth at different values for the specific energy (which is set by the throttle
position), a 𝑉2 − 𝑦2 vs. 𝐸 diagram was computed for 𝐵 = 0.5 m, see Figure 3.3. As can be
seen, the flow velocity and flow depth are dependent on the specific energy, which ranges
between 𝐸𝑚𝑖𝑛 and 𝐸𝑚𝑎𝑥 . Note that 𝐸𝑚𝑖𝑛 corresponds to the value of the specific energy at
the intersection of the 𝐵 = 0.5 m line and the Fr = 1 line in Figure 3.2. In fact, if the specific
energy would be decreased below 𝐸𝑚𝑖𝑛 , the upstream flow would become subcritical; a
condition which does not allow for the flow depth to increase during a channel transition.
Based on Figure 3.3 it can be stated that for a channel width of 𝐵 = 0.5 m, a hydrokinetic

16RET1012
22
5.5 0.48

5.0 0.45

4.5 0.42

4.0 0.39
V2 [m/s]

y2 [m]
3.5 0.36

Emax
Emin

3.0 0.33

2.5 0.30

2.0 0.27
0.5 0.7 0.9 1.1 1.3 1.5 1.7 1.9
E [m]

Water velocity [m/s] Water level [m]

Figure 3.3 Variation of 𝑉2 and 𝑦2 with specific energy for 𝐵 = 0.5 m. Decreasing the specific energy below
𝐸min would not allow for a raise of the flow depth with respect to the upstream flow.

turbine can be tested at water velocities, 𝑉2 , ranging from approximately 2 - 5.4 m/s and
corresponding flow depths, 𝑦2 , ranging from approximately 43 – 30 cm.

The channel transition from 𝐵1 = 1 m to 𝐵2 = 0.5 m is considered to provide a great


improvement with respect to the flow conditions at which a hydrokinetic turbine can be
tested in the water flume in the hydrological laboratory of the AdeKUS. Hence, said
transition will be part of the modifications that are suggested to be made to the test facility
in order to conduct experimental research on hydrokinetic turbines. In the present work
only the basic functional geometry of the transition region is provided, i.e., supporting
structures, auxiliary parts and material choices are not discussed herein.

In order to avoid large velocity gradients, a gradual contraction of the flume is proposed,
followed by a straight test section and a gradual expansion to the original flume width,
respectively. The shape of the contraction and expansion zone are arbitrarily modeled
with a 5th order polynomial, optimized to ensure tangential propagation, i.e.,

𝑧(𝑥 ′ ) = 𝐵1 (𝐵1 − 𝐵2 ) ∙ [6(𝑥 ′ )5 − 15(𝑥 ′ )4 + 10(𝑥 ′ )3 ] (3.6)


In equation (3.6), 𝑥 ′ represents a distance in the flow direction, normalized by the
transition length, 𝐿, whereas z denotes the width of the transition region.

Taking into account: i) the dimensions of commercially available sheet metal, which will
most probably be the material of choice for the sidewalls of the test section and transition
regions, and ii) typical test section lengths in similar (hydro/air)foil experiments, the

16RET1012
23
transition length and the test section length of the experimental setup at hand are
conveniently set at 0.9 m and 1.2 m, respectively. Substitution of 𝐵1 = 1 m and 𝐵2 = 0.5 m
into equation (3.6) leads to

𝑧(𝑥 ′ ) = 3(𝑥 ′ )5 − 7.5(𝑥 ′ )4 + 5(𝑥 ′ )3 (3.7)


The shape of the test section with its adjacent transition regions is shown in Figure 3.4 and
Figure 3.5. Given the fact that testing of the hydrokinetic turbines should ideally be
performed in a homogenous flow, it is important to at least provide an estimate of the
boundary layer thickness in the test section. However, due to the fairly complex shape of
the flume transition, the boundary layer cannot be easily studied from an analytical
perspective, using for example the Blasius solution for boundary layer flow. Instead, a
numerical study was carried out, which provides more insights in this regard (see Section
3.2.2). Another important aspect to consider, is the blockage effect, which is the distortion
of the flow whenever an obstruction, such as a fuselage (or nacelle) is introduced therein.
In fact, the obstruction prevents the formation of proper vena contracta, influencing the
acceleration (or deceleration) of the flow, and consequently, in the case of a lift-based
horizontal axis turbine, the power coefficient. Additionally, the free surface proximity of
the HK-turbine also affects the blockage effect due to free surface deformation
(asymmetrical wake), as is shown in Ref. (Kolekar, et al., 2019). Quantification of the
impact of blockage effects demands for an incorporation of the flow around the
obstruction into the wake model or experimental flow field investigations to characterize
wake structures, vortex structures, and flow in bypass regions; this is beyond the scope of
this work. Typically, in wind tunnel experiments the ratio of obstructed area to the total
cross sectional area of the wind tunnel (blockage ratio) is less than 10% (Chen & Liou,
2011; Hansen, 2008). This ratio is adopted herein (as a precursor to an in-depth study on
blockage effects), as there are no guidelines found in literature on blockage ratios in water
flume experiments.

0.6
Contraction zone Test section Expansion zone
0.4

0.2
𝐵2 = 0.5 m
z* [m]

0
𝑧

-0.2

-0.4

-0.6
0 0.5 1 1.5 2 2.5 3
x* [m]

Figure 3.4 Shape of the contraction, expansion and test section, expressed in Cartesian coordinates, (𝑥 ∗ , 𝑧 ∗ ).

16RET1012
24
In an attempt to eliminate large vortices (or Eddies) and transverse velocity components
in general from the upstream flow2, a honeycomb-like flow straightener built from PVC
pipes is proposed. PVC pipes are commonly available in hardware stores and are
inexpensive. It should be noted, however, that the non-zero thickness of the pipes will
cause small vortices at the pipe-exit, and it is therefore proposed to chamfer the ends of
the PVC pipes using a chamfering tool, to reduce said effect. Notwithstanding these
measures, small vortices which will inevitably be present in the flow; their significance on
the projected experiments is not studied herein. A literature study has not provided much
clarity on this matter, however, within this context, a study (Mikkelsen, 2013) has been
found which shows that free stream turbulence has a relatively small effect on the
performance characteristics of a wind turbine, and that higher levels of turbulence actually
leads to faster recovery of the wake downstream of the turbine due to increased mixing.
Eventhough the power extraction is slightly reduced with free stream turbulence, it seems
like the effect on the recovery of the wake is larger, which leads to a netto higher power
extraction.

In principle, for a given diameter, the length of the PVC pipes must be chosen such that a
fully developed flow can be achieved within; the travelled distance of flow before
becoming fully developed is referred to as the entrance length. For a pipe the entrance

Figure 3.5 Isometric view of the proposed flume transition.

length, 𝐿𝑒 , is correlated to the Reynolds number of the entering flow as follows (White,
2015):
𝐿𝑒 1
= 4.4 ∙ 𝑅𝑒 ⁄6 (3.8)
𝑑𝑝𝑖𝑝𝑒
In equation (3.8) 𝑑𝑝𝑖𝑝𝑒 equals the (inner) diameter of the pipe. It should be noted that this
correlation is only valid for 𝑅𝑒 > 4000. The Reynolds number for a pipe flow is computed
using
𝜌𝑢𝑑𝑝𝑖𝑝𝑒 𝑢𝑑𝑝𝑖𝑝𝑒
𝑅𝑒 = = (3.9)
𝜇 𝜈

2As was previously mentioned, a wavy flow or a flow with large vortices may constitute to a non-
homogenous and unbalanced pressure distribution on the turbine rotor, which is not desirable.

25
16RET1012
where 𝜌, 𝜇 and 𝜈 represent the density, the dynamic viscosity and the kinematic viscosity
of the fluid, respectively, and 𝑢 is the flow velocity within the pipe.

Considering the limiting case, whereat 𝑢 = 5.6 m⁄s and selecting 2” PVC pipes for the
straightener, the maximum Reynolds number of the flow entering the pipes was
calculated to be 285 ∙ 103 (with 𝜈 = 10−6 𝑚2 ⁄𝑠 ) and therefore the entrance length, 𝐿𝑒 ,
must be 1.81 m, as computed with equation (3.8). In an attempt to estimate the pressure
loss across the pipes, the Hagen-Poiseuille equation (see equation (3.10)) was employed,
resulting in ∆𝑝 = 22429 Pa = 2.29 mH2 O (with 𝜇 = 8.90 ∙ 10−4 Pa ∙ s).
8𝜋𝜇𝐿𝑄
∆𝑝 = (3.10)
𝐴2pipe

Figure 3.6 provides a graphical representation of the flow straightener composed from
PVC pipes which are stacked on top of each other along the width of the flume. Note that
at least four rows of pipes are necessary given the water level of the upstream flow. The
flow straightener is to be placed exactly in front of the contraction region.

Figure 3.6 Isometric view of the proposed flow straightener composed from 2" PVC pipes.

As was pointed out earlier in this section, currently water is discharged almost
perpendicularly into the flume via two pipes from the top reservoir of the testing facility
(see Figure 3.1). Obviously this configuration causes a lot of turbulence (and kinetic energy
losses) and therefore a modification, which is fairly simple to implement, is proposed; both
pipes originating from the top reservoir can be extended and repositioned slightly, with
90-degree elbow fittings connected at their end, such that an almost tangent propagation
of the pipe flow into the flume is achieved. In Figure 3.7 a depiction of the latter is given.

16RET1012
26
Figure 3.7 Isometric view of the modified flume water feeding configuration.

3.2.2 CFD simulation of the (open) channel flow


In the present section results from a CFD simulation of the flow in the water flume testing
facility is presented. All of the modifications that have been proposed previously are
considered in this simulation, whose goal was to provide more accurate and insightful
results with respect to the earlier (preliminary) calculations on the velocity and the water
level in the test section for the hydrokinetic turbine. However, due to computational
limitations, said goal could only partly be realized. In fact, the obtained results must be
considered to provide a qualitative picture of the flow behavior in the flume; from a
quantitative perspective, on the other hand, no reliability can be guaranteed. Nevertheless,
the simulation results are consistent with the previous calculations, which provides some
confidence that the computed numbers are within a reasonable range. Taking into account
the foregoing, the CFD simulation performed as part of this work must be regarded as a
first attempt, which still needs a follow-up.

Particularly the velocity vector field in the test section for the hydrokinetic turbine was
scrutinized. For this purpose, a scenario whereby no throttling in the discharge pipes from
the top reservoir is performed was considered, consistent with the rationale discussed
earlier. Important aspects regarding the built CFD model, including boundary conditions,
assumptions, configurations and properties, are listed below:

 Given the rather large dimensions of the computational domain and calculation
limits imposed by the available computer and simulation software (ANSYS CFX),
a subdivision into three independent models was necessary, namely:
- Model 1: The channel water feeding pipes (see Figure 3.7) and the channel
itself up to the flow straightener.
- Model 2: The flow straightener (see Figure 3.6).
- Model 3: The HK turbine test section, including its adjacent transition
regions (see Figure 3.5).
Starting with the first model, a CFD simulation was performed, from which
computational results at its outflow boundary were exported into the next model
to perform the subsequent simulation.
 For Model 1 the hex-dominant mesh method with a consistent mesh size of 0.05 m
throughout the entire computational domain was employed. The hex-dominant

16RET1012
27
method was selected because it has shown to be a very suitable choice for large
volume bodies. Simulation attempts with a larger mesh sizes (even with mesh
refinement at locations where large velocity gradients were expected) were
unsuccessful; poor grid convergence and void spaces in the computational domain
were observed, deviating from what is physically meaningful.
 For Model 2 the tetrahedrons mesh method was used with an automatically
generated mesh size of 1.5 ∙ 10−3 m, as the hex-dominant mesh method generated
poorly shaped hexes due to the relatively small size of the flow straightener pipes.
 For Model 3 the hex-dominant mesh method with a uniform mesh size of 0.05 m,
was employed. Due to computational limits the mesh size could not be reduced
any further, which in hindsight was not necessary, as no numerical problems were
encountered, results were qualitatively as expected and did not deviate
significantly with simulations conducted using larger mesh sizes. In fact, the
residual RMS values for the internal loops (i.e., at each time step) were less
than 10−4 , which shows grid convergence. Moreover, after approximately 5
minutes, a steady-state condition (considering the monitored variables: velocity
and water level) was reached.
 The analysis type in ANSYS CFX was chosen to be transient. It should be noted
that, at first, a steady-state simulation was attempted which, however, failed to
converge; although the flow is inherently transient, in a statistical sense one may
assume statistical stationarity. While the failure to converge may be due to poor
mesh quality affecting the solution and/or dissipative errors in the solution
discretization (or even incorrect solution parameters), the unsteadiness in the flow
might lead to simulation problems as well. In fact, it is known that flows involving
the presence and interaction of two or more fluids, separated by sharply defined
interfaces, such as in the free-surface problem at hand, are best predicted with a
transient simulation, from which time-averaged statistics may be obtained
afterwards.
 Considering that the water flow is expected to be turbulent in some regions of the
computational domain(s), the standard k-𝜀 turbulence model was activated. The
(standard) k- 𝜀 turbulence model is widely used to simulate mean flow
characteristics for turbulent flow conditions, mainly because of its simplicity (in
terms of input parameters) and its broad application envelope. It does however
perform poorly when large adverse pressure gradients are present, as opposed to
the k-𝜔 turbulence model or realizable k-𝜀 turbulence model, for example. In the
present case, adverse pressure gradients are absent and therefore not an issue.
 A no-slip condition is assumed in all three models, i.e., the water velocity at all
fluid-solid interfaces is equal to that of the solid boundary.
 The interaction between (ambient) air and the water in the flume is simulated by
employing the Volume of Fluid (VoF) model and the surface tension model. In
these models water and air were configured as the primary and secondary phase,
respectively, with a water volume fraction of one at the inlet and zero at the outlet
(initially; with time a mixture develops). The surface tension coefficient was set to
0.072 N⁄m (which is valid for water and air as the primary and secondary fluids,
respectively, at atmospheric conditions). The occurrence of backflow at the outlet
was limited to air only.

16RET1012
28
 The CFD models were solved using a second-order backward Euler method with
a time step size of 0.1 s, given the transient nature of the simulation. Assuming a
one-dimensional advection dominated flow, the Courant-Friedrichs-Lewy (CFL)
convergence condition is satisfied with the chosen physical time-step size (i.e., the
Courant number is less than unity for all computational domains).
 A uniformly distributed water velocity of 5.6 m⁄s (≜ 𝐸max ) was assigned to the
water inlet of the pipes as shown in Figure 3.7.
 The channel transition section was fixed at the end of the water flume, as it can be
argued that the water velocity profile in the flume is most stable at this location
(with respect to other locations closer to the water feed).

Obviously quite some interesting results can be obtained from the CFD simulations that
were conducted, however, i) as stated previously, the obtained results are questionable as
no grid dependency study could be conducted due to computational limitations (note that
larger mesh sizes resulted into diverging solutions, whereas a smaller mesh sizes were not
possible due to computational limitations), ii) with respect to the design of a scaled model
of the HK-10 turbine, the results in the test section (Model 3) are by far the most important,
as the accumulated effect of all adaptations is exhibited herein, and therefore only the
numerical results of Model 3 will be discussed. Regarding the (exact) influence of the
modified water feeding pipes and the flow straightener on the results in the test section,
the following can be stated: from a qualitative point of view, these modifications will, as
far as can be argued, positively affect testing conditions for the hydrokinetic turbine,
which was already considered a decisive factor for these modifications to serve as a basis
for the continuation of this research project. It is however recommended to quantify their
influence and eventually optimize their shape/configuration, prior to actual
implementation in the testing facility.

Computations have shown that the flow upstream of the channel contraction (coming out
of the flow straightener) is supercritical, has an average free-stream velocity of 5.1 m⁄s
and a water level of approximately 14 cm, see Figure 3.8a. Due to the contraction, a
hydraulic jump occurs, which causes the water level to increase significantly as was
predicted earlier. Figure 3.8b shows a contour plot of the water volume fraction in the mid
plane of the test section (perpendicular to the flow direction), from which the water
distribution in this computational domain becomes apparent; the water reaches a level of
approximately 28 cm, disregarding the elevated sides. Noteworthy, this result is in good
accordance with the preliminary calculations presented previously (see Figure 3.3; at 𝐸 =
𝐸max the water level was computed to be 29.7 cm). What is remarkable, however, is the
extensive diffusion of water at the flow surface; it is suspected that this phenomenon can
be observed from the numerical solution due to a rather large mesh size (and not due the
physical circumstances).

The velocity in the mid plane of the test section is visualized by means of contour plots in
Figures 3.9, 3.10 and 3.11. As can be seen from Figure 3.9, the free-stream velocity of the
flow in x-direction (parallel to the flume) is about 4.6 m/s.

16RET1012
29
(a)

(b)

Figure 3.8 (a) Water volume fraction in the plane just before the flow contraction and (b) in the mid plane of
the test section; the width transition from 𝐵1 = 1 m to 𝐵2 = 0.5 m induces a hydraulic jump causing the water
level to increase significantly.

In the transverse directions, water velocities are relatively small as can be seen from
Figures 3.10 and 3.11, particularly when considering the area of interest, which is the
region below the water surface (refer to Figure 3.8b). Interestingly, in the horizontal
direction (z-direction) an asymmetric velocity profile is observed. This asymmetry is a
result of the forced flow contraction, which induces horizontal velocity components in
opposing directions.

16RET1012
30
Figure 3.9 Velocity distribution in the mid plane of the test section for velocities parallel to the flume (x-
direction).

Figure 3.10 Velocity distribution in the mid plane of the test section for velocities in the horizontal direction
(z-direction), perpendicular to the main flow direction.

Contrarily to the comparison made previously between the preliminary calculations and
the CFD results regarding the water level in the test section, a significant difference exists,
however, for the water velocity when comparing both computation methods. When
inspecting Figure 3.3, it can be seen that at 𝐸 = 𝐸max the water velocity in the test section
was calculated to be 5.4 m/s, whereas the CFD computations predict a velocity which is
significantly lower, see Figure 3.9. This difference can be attributed to the fact that the
energy conservation method (based on the assumption that the specific energy of the flow
remain constant throughout the length of the flume) did not account for kinetic energy
losses, which explains why the CFD computation outputs lower velocities. It is however
remarkable that although there is a significant difference between the methods with

16RET1012
31
Figure 3.11 Velocity distribution in the mid plane of the test section for velocities in the vertical direction (y-
direction), perpendicular to the main flow direction.

respect to the water velocity, the water level in the test section compares quite well, as
stated previously. Scrutinizing Figure 3.3, provides insightful findings in this regard; the
water level appears to be a weaker function of the specific energy close to 𝐸 = 𝐸max ,
compared to the water velocity, i.e., kinetic energy losses have a greater effect (relatively)
on the water velocity than on the water level.

Based on the results from this section, it can be conservatively concluded (considering the
previously mentioned shortcomings) that the suggested adaptations improve the flow
conditions (particularly with respect to the water level) such, that testing of hydrokinetic
turbines in the hydrological facility at the AdeKUS can be made possible. In fact, with no
throttling applied in the discharge pipes from the top reservoir in the testing facility, water
reaches a water level of roughly 30 cm in the test section, with a velocity of 4.6 m⁄s (other
scenarios are obviously possible, but arguably less ideal). This velocity is almost
homogenously distributed throughout an area that is considered sufficiently large to test
hydrokinetic turbines with practical dimensions. Based on a blockage ratio of 10% (this
was discussed earlier) and the computed water level in the test section, the diameter of
the turbine’s rotor should not exceed 16 cm.

16RET1012
32
4 BEM SIMULATION RESULTS
4.1 Simulation method
All computations involved in the BEM analysis have been performed using MATLAB®
R2014a, a product from MathWorks®. Equations presented in Section 2 have been
implemented in several script files and function m-files, using different programming
techniques. As far as non-linear equations were involved, use has been made of the built-
in numerical solver fsolve, based on the Trust-Region-Dogleg Algorithm to find
solutions. However, for ill-posed problems the bisection method (this is a very robust
numerical method) was employed due to convergence issues encountered with fsolve.
Note that the numerical results obtained have been exported to Microsoft Excel 2013 for
visualization in graphs, as this was found to be more convenient with respect to built-in
graphing in MATLAB®.

Drag and lift coefficients, which are important quantities involved in the BEM analysis to
compute hydrodynamic forces (refer to Section 2.5), have been obtained using XFOIL
(version 6.99) via a COM interface with MATLAB®. XFOIL is an interactive program for
the design and analysis of subsonic isolated airfoils developed at the Massachusetts
Institute of Technology (MIT). Basically, given coordinates specifying the shape of an
airfoil (which is the NACA 4415 in this project), the Reynolds number and Mach number,
XFOIL can compute drag and lift characteristics.

4.2 Computational procedure, results and discussion


In this section the computational procedure and the obtained BEM simulation results are
discussed for both the prototype rotor and the scaled model rotor (𝐷mod = 0.16 m, see
previous section). Said rotors are respectively simulated in water and air flows. Based on
results from this section a scaling factor was determined via a dimension analysis in order
to transpose obtained wind tunnel experimental results to prototype results (see Section
6).

In order to compute hydrodynamic forces and ultimately determine the power curve for
a turbine, it is necessary to first define the geometry of its rotor, which is optimized based
on the nominal operating conditions of the turbine. In the present case, these conditions
have been set by the University of Brasilia (UnB) and the Federal University of Para
(UFPA) for the Hydro-K project, which was already mentioned in Section 1. In fact, based
on these conditions (𝑛 = 35 rpm, 𝑢∞ = 2.5 m⁄s and 𝐷 = 2.1 m , see also Table 2.1) the
geometry of the HK-10 turbine was established.

For validation purposes and as part of the procedure to compute scaling factors (this will
be discussed in depth in Section 6.1.2) required for experimental analysis with physical
scale models tested in flow conditions dissimilar from design conditions, part of the
computations carried out by researchers at the University of Brasilia have been repeated.
Note that validation was deemed necessary because i) errors were found in some key
equations documented in their work (Brasil & Lavaquial, 2017), whereby it was unclear if
these were just typographical errors or incorrect implementation of theory in their model,
and ii) an inconsistency between the documentation and their BEM model was discovered
regarding the use of a cascade flow correction factor, which led to reasonable doubt.
Moreover, their implementation of the BEM method (which was also done in MATLAB®)

16RET1012
33
was considered very confusing and due to poor programming annotation it was
practically impossible to understand its functioning. As a result, all BEM computations
required for the project described herein have been carried out independently from the
model developed by the Brazilian researchers.

In order to compute the optimal flow conditions for the HK-10 turbine and its optimal
power coefficient, 𝐶𝑃,opt, the following steps have been carried out:

1. Establish the nominal operating conditions, i.e., the rotational speed, 𝜔 , the
(upstream) flow speed, 𝑢∞ , and the rotor diameter, 𝐷;
2. Divide the rotor blade(s) into N sections, each at a distance 𝑟𝑖 (𝑖 = 1: N) from the
rotor’s axis of rotation (in this work N was set to 100);
3. Compute 𝜆𝑟,𝑖 = 𝜔𝑟𝑖 for 𝑖 = 1: N;
4. Compute optimal axial induction factors, 𝑎𝑖 , for each blade element (𝑖 = 1: N) by
solving equation (2.34) for 𝑎;
5. Compute optimal angular induction factors, 𝑎𝑖′ , for each blade element (𝑖 = 1: N)
by solving equation (2.33) for 𝑎′;
6. Calculate optimal entrance angle, 𝜙𝑖 , for each blade element ( 𝑖 = 1: N ) using
equation (2.23);
7. Compute the (optimal) power coefficient, 𝐶𝑃,opt for the nominal operating
condition using equation (2.29).

The optimal (or maximum) power coefficient, 𝐶𝑃,opt for the HK-10 turbine was found to
be approximately 0.48 for TSR = 1.61 (see Figure 4.1), conforming to the nominal
operating condition. Note that this value was calculated without considering Prandtl
losses and therefore the actual maximum is lower, which will be shown at a later point in
this section.

0.7

0.6 Betz limit = 16/27 ≈ 0.593

0.5 𝐶𝑃 ≈ 0.48
CP, opt [-]

0.4
TSR = TSR nom = 1.61

0.3

0.2

0.1

0.0
0 1 2 3 4 5 6 7
TSRnom [-]

Figure 4.1 Maximum 𝐶𝑃 value corresponding to the nominal operating condition, disregarding Prandlt losses.

16RET1012
34
Next, the optimized geometry of the HK-10 turbine was determined using input from the
previous procedure and XFOIL simulation data in order to compute chord lengths, 𝑐𝑖 , and
blade pitch angles, 𝜃𝑖 . More specifically, XFOIL was employed to calculate the hydrofoil’s
(NACA 4415) optimal angle of attack, 𝛼𝑜𝑝𝑡 , and its corresponding drag and lift coefficients
(resp. 𝐶𝐿 (𝛼) and 𝐶𝐷 (𝛼)). Note that 𝛼𝑜𝑝𝑡 is defined as the angle of attack for which the ratio
between 𝐶𝐿 and 𝐶𝐷 is maximal.

Due to poor convergence of XFOIL for some values of 𝛼 under simulated conditions a
function was fitted to converged data points to allow for interpolation and to set up a
method that guarantees convergence during BEM calculations. Herein the same fitting
function as proposed in Ref. (Brasil & Lavaquial, 2017) for this purpose, was employed,
i.e.,
5

𝐶𝐿 (𝛼) = ∑ 𝑘𝑖 [ln(𝛼 + 10°)]𝑖 (4.1)


𝑖=0
and
5

𝐶𝐷 (𝛼) = ∑ 𝑝𝑖 [ln(𝛼 + 10°)]𝑖 (4.2)


𝑖=0

where 𝑘𝑖=1…5 and 𝑝𝑖=1…5 are fitting parameters.

Figure 4.2 and Figure 4.3 show the results from the performed regression analysis for
respectively the prototype (under design conditions, Reprot = 1.1 ∙ 106 ) and the test model
(under wind tunnel testing conditions, Remod = 3.0 ∙ 104 , see Section 6). Moreover, in
Figure 4.4 the ratio between 𝐶𝐿 (𝛼) and 𝐶𝐷 (𝛼) is visualized graphically for the prototype,
from which it is evident that a maximum exist at = 6.5° , which is referred to as 𝛼𝑜𝑝𝑡 , as
was previously indicated.

1.8 0.30

1.6
0.25
1.4

1.2 0.20
CD [-]

1.0
CL [-]

0.15
0.8

0.6 0.10

0.4
0.05
0.2

0.0 0.00
-5 -1 3 7 11 15 19 23 27 31
α [˚]

Lift Coefficient, CL (X-Foil) Lift Coefficient, CL (fitted)

Drag Coefficient, CD (X-Foil) Drag Coefficient, CD (fitted)

Figure 4.2 Lift and drag coefficients at design conditions for different angles of attack within the simulation
range.

16RET1012
35
1.4 0.40
1.2 0.35
1.0
0.30
0.8
0.6 0.25

CD [-]
CL [-]

0.4 0.20
0.2 0.15
0.0
0.10
-0.2
-0.4 0.05

-0.6 0.00
-5 -1 3 7 11 15 19 23 27 31
α [˚]

Lift Coefficient, CL (fitted) Lift Coefficient, CL (X-Foil)

Drag Coefficient, CD (fitted) Drag Coefficient, CD (X-Foil)

Figure 4.3 Lift and drag coefficients at wind tunnel testing conditions for different angles of attack within the
simulation range

140

120

100
𝛼 = 𝛼opt = 6.5°
CL/CD [-]

80

60

40

20

0
-5 -1 3 7 11 15 19 23 27 31
α [˚]

Figure 4.4 The optimal angle of attack for the NACA 4415 foil is approximately 6.5. The maximum was
determined from a cubic spline fit.

The regression lines in Figure 4.2 and Figure 4.3 can be considered good fits from a
quantitative point of view. There are, however, some curvatures (namely at low values
of 𝛼, see for example Figure 4.2 between 𝛼 = −5° and 𝛼 = 7°) that cannot be observed
from the XFOIL simulation data and therefore from a qualitative perspective the

16RET1012
36
performance of the regression functions (4.1) and (4.2) is rather poor. Nevertheless, since
the BEM analysis does not involve second (or third) order quantities the fitted functions
are unarguably suitable in the present case.

With the drag and lift characteristics obtained through XFOIL, the optimal blade geometry
was determined using the following procedure:

1. Consider previously calculated values: 𝑎𝑖=1…𝑁 and 𝜙𝑖=1…𝑁 ;


2. Determine 𝛼𝑜𝑝𝑡 (= 6.5°) from Figure 4.4 and compute 𝐶𝐿 (𝛼𝑜𝑝𝑡 ) and 𝐶𝐷 (𝛼𝑜𝑝𝑡 ) from
equations (4.1) and (4.2);
3. Calculate the pitch angles, 𝜃𝑖 , for all blade elements (𝑖 = 1: N) using 𝜃𝑖 = 𝜙𝑖 − 𝛼𝑖
(see also Figure 2.8) with 𝛼𝑖 = 𝛼𝑜𝑝𝑡 ;
4. Compute the chord length, 𝑐𝑖 , for each blade element (𝑖 = 1: N) with equation
(2.64).

The optimal blade geometry of the HK-10 rotor in terms of pitch angles and chord lengths
is shown in Figure 4.5. Note that the chord length and the distance to each blade element,
𝑟, have been normalized for convenient comparison with other rotors found in literature.

0.45 50

0.40 45
Normalized Chord Length, c/R [-]

40
0.35

rotor tip position


rotor hub position

35
0.30

Pitch Angle [˚]


30
0.25
25
0.20
20
0.15
15
0.10
10

0.05 5

0.00 0
0.0 0.2 0.4 0.6 0.8 1.0 1.2
Normalized Radius, r/R [-]

Normalized Chord Length Pitch Angle

Figure 4.5 Optimized rotor geometry in terms of pitch angles and chord length.

With the geometry of the HK-10 rotor now defined, the power curves for both the
prototype and test model can be computed using BEM theory. The computation steps are
summarized as follows:

1. Consider previously calculated data: 𝑐𝑖=1:N and 𝜃𝑖=1:N ;


2. Provided a fixed upstream flow velocity, 𝑢∞ , compute for a range of different

16RET1012
37
rotational velocities, 𝜔, for each blade element the axial and angular induction
factors, 𝑎 and 𝑎′, using an iterative procedure:
i. Provide an initial guess for 𝑎 and 𝑎′ as follows: 𝑎 = 1⁄3 and 𝑎′ = 2⁄(9𝜆𝑟,𝑖 );
ii. Compute 𝜙 with equation (2.23);
iii. Compute the angle of attack with 𝛼 = 𝜙 − 𝜃𝑖 ;
iv. Calculate Prandtl’s correction factor, 𝐹, using equations (2.57), (2.58) and
(2.59);
v. Using XFOIL compute the drag and lift coefficients, 𝐶𝐷 (𝛼) and 𝐶𝐿 (𝛼) ,
followed by 𝐶𝑁 and 𝐶𝑇 using equations (2.45) and (2.46), respectively;
vi. Compute 𝑎 and 𝑎′ with equations (2.60) and (2.61), respectively, ;
vii. If the computed axial induction factor, 𝑎𝑖 , is larger than 𝑎𝑐𝑟𝑖𝑡 = 0.45, use
Buhl’s empirical relation (equation (2.62)) instead of equation (2.60) for
subsequent iterations;
viii. Compare the actual values for 𝑎 and 𝑎′ with those from the previous
iteration. If the difference is less than the required accuracy (which was set
to 10−8) terminate the loop, else repeat steps (ii) – (viii);
3. For all rotational velocities considered, with their corresponding set of axial and
angular induction factors for all blade elements, compute the power coefficients
by numerically integrating equation (2.29).

As mentioned previously, the numerical discretization was done by dividing the blades
into 100 sections (𝑁 = 100) of equal length, ∆𝑟, between the hub and the tip of the rotor.
Noteworthy, the simulation results do not differ significantly from simulations with far
less sections (the simulation has also been performed with 𝑁 = {10, 25, 50}; these results
are not shown herein) and therefore the results are considered independent from the
numerical discretization. All results obtained have a numerical accuracy of at least 10−6
(refer to Section 4.1 for details on the numerical methods employed).

In Figure 4.6 the results from the computations according to the procedure just presented
is shown. Note that the BEM simulations were conducted at flow speeds of 2.5 m/s (water
velocity) and 15 m/s 3 (wind velocity), for the prototype rotor and scaled model rotor,
respectively. Rotational speeds were varied between 10 and 75 RPM for the prototype and
1000 and 5500 RPM for the scaled model. These ranges were determined based on a trial
and error approach with the goal to generate a clear view of the power curves.

From the results it is evident that the power coefficient, 𝐶𝑃 , of the HK-10 rotor reaches a
maximum of approximately 0.38. Recall from Figure 4.1 that the maximum power
coefficient was previously estimated to be 0.48, a value that was, however, determined
without taking into consideration Prandtl losses. It can therefore be concluded that Prandtl
losses are quite significant for the HK-10 turbine. For the scaled model of the HK-10 rotor
the maximum power coefficient was computed to be 0.25 as can be seen from Figure 4.6.

3This value corresponds to wind speed at which the scaled model of the HK-10 rotor was tested in
the wind tunnel in Brasilia, see Section 6.1.1.

38
16RET1012
0.45

0.40

0.35

0.30

0.25
Cp [-]

0.20

0.15

0.10

0.05

0.00
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5
TSR [-]

Power curve (prototype) Power curve (model)

Figure 4.6 The power curves for both the prototype and the scaled model of the HK-10 rotor. Note that the
power curve for the lab-scale model is computed for wind tunnel testing conditions, as will be discussed in
Section 0.

Remarkably, for both the prototype and the scaled model, the top of the power curve is
located at a tip-speed-ratio of around 2.2, which is different from the design value
(TSR nom = 1.61). Although, no definite explanation for this phenomenon was found, it
issuspected that XFOIL does not correctly predict lift and drag coefficients at higher tip-
speed-ratios for the NACA 4415 profile at the simulated Reynolds numbers. The validity
of this hypothesis should be confirmed by means of dedicated experiments and/or CFD
simulations, which are beyond the scope of the present work. However, anticipating the
results from the next section, it can be pointed out that the experimentally determined
power curve for the HK-10 rotor does have its maximum value located around TSR = 1.61,
which, in principle, already narrows the problem down to the BEM model (and not the
actuator disc model employed for the optimization of the rotor geometry).

16RET1012
39
16RET1012
40
5 DESIGN AND MANUFACTURING OF A LAB-SCALE HK-10
TURBINE
5.1 Design and manufacturing of physical components
A lab-scale model of the HK-10 turbine equipped with torque and RPM sensors was
already developed and tested in a wind tunnel at the University of Brasilia. As stated in
Section 1.2, a wind tunnel was employed due to the absence of a water flume testing
facility. Although some very useful data can be obtained from wind tunnel experiments,
particularly under steady-state conditions, some limitations have been identified (again,
refer to Section 1.2) which demand for actual underwater experimenting with the HK-10
turbine. The lab-scale model developed by the researchers at the University of Brasilia
performs demonstrably well (from a functional point of view) in a wind tunnel setup, but
due its open generator (and measuring devices) enclosure is not suitable for underwater
experiments. Moreover, the following shortcomings/issues require for a modified version
of the lab-scale turbine: i) considering experimental testing at the hydrological laboratory
at the AdeKUS, the diameter of the rotor should not exceed 16 cm, as was determined in
Section 3.2; the existing scale model was equipped with a 22 cm rotor, therefore being too
large, ii) the turbine housing has a diameter which is inevitably larger than the rotor hub,
still a smooth transition of the flow must be guaranteed in order to avoid large pressure
gradients and turbulence immediately behind the rotor. Such can be achieved by means
of a static flow deflector, which is, however, lacking in the existing scale-model, iii) the
torque measuring device and driveline components arguably provide (based on BEM
results in the previous section) the required structural integrity when testing the HK
turbine in a water flow such as predicted in Section 4.2.

A modified design, addressing the mentioned issues was developed, as is shown in Figure
5.1. It should be noted that some design choices were made based on the available
materials in Brasilia4, as the lab-scale model had to be manufactured thither. A description
of this design is provided next (note that the numbering corresponds to the labeling in
Figure 5.1):

1. Lab-scale HK-10 rotor: A four bladed horizontal axis lift based rotor with the
NACA 4415 profile. Note that the lab-scale model is geometrically similar to the
prototype (HK-10 rotor, see also Figure 4.5), i.e., the same ratio of linear scale exist
for all coordinates describing the shape of the respective rotors. Following the
outcomes from Section 0, the diameter of the lab-scale rotor is chosen to be 16 cm;
this corresponds to a scaling factor of exactly 13.125 with respect to the prototype.
The material employed to manufacture this part (with 3D printing technology) is
acrylonitrile butadiene styrene (ABS) as it has high tensile strength and is very
resistant to physical impacts. Moreover, ABS can be sanded and machined
relatively easy with respect to other common 3D printing materials such as PLA,
PET and nylon, which is an advantage given the followed manufacturing
procedure.

4Part of the research project documented herein was conducted at the University of Brasilia, Brasil,
during a period of 2.5 months.

41
16RET1012
12 13
9 3
8
1
4 2

11

10

6
5

14

Figure 5.1 Section view of the modified design of the lab-scale turbine suitable for underwater experiments.
The numbering in the figure corresponds to the description in the text.

2. Flow deflector: In order to avoid dynamic pressure build up behind the rotor and
allow for a smooth development of the flow, a flow deflector is introduced
following the shape of a polynomial, optimized such that there is tangent relation
between the deflector and the rotor hub and generator housing, respectively. The
flow deflector is manufactured with ABS as well and slides onto the bearing block
described next.
3. Bearing block: An aluminum bearing block houses two roller type bearings on
both ends ensuring constrained movement of the intermediate shaft (see part
number 5) with minimal friction involved. It is extremely important that no lateral
movement of the shaft is allowed, in order to avoid bending moments in the torque
sensor (part number 10); the existing lab-scale model was lacking this feature. The
bearing block is fixed by means of countersunk hex bolts to the generator housing
(part number 12).
4. Roller bearings (2x): Deep groove ball bearings from SKF (W63801-2RS1). These
are sealed type stainless steel bearings, with dynamic and static load ratings
exceeding the minimal requirements, and therefore suitable for this application.
The inner and outer diameter of these bearings are 12 mm and 21 mm ,
respectively, whereas the width is 7 mm. It should be noted that smaller bearings,
with the required characteristics, were not available in Brasilia. Both bearings are
press-fitted into the bearing block (interference fit, 0.02 mm) with an additional

16RET1012
42
cylindrical retaining compound (LOCTITE® 638) to fill eventual imperfection in
the mating surface.
5. Intermediate shaft: A steel shaft connects the rotor with the generator shaft via a
semi-rigid coupling (part number 6). The diameter of the intermediate shaft
corresponds to the inner diameter of the mentioned bearings (= 12 mm). Using the
same retaining compound as mentioned previously, the shaft is secured into
position within the inner bearing rings.
6. Shaft coupler: A brass coupling which is secured onto the generator shaft (𝑑 =
5 mm) by means of a bolt lock. The intermediate shaft slides into this coupling
consisting of both a cylindrical and flat part for concentric positioning and torque
transmission, respectively. A thin gap is designed between the intermediate shaft
and the coupling, which is filled with RTV silicone upon assembly to reduce
vibration as a consequence of possible misalignment.
7. Front O-ring seal: The bearing block partially slides into the turbine housing,
potentially exposing the generator and measuring devices inside to water
contamination via the gap (= 0.05 mm) existing between the concentric surfaces.
To prevent water from leaking into the generator housing, a square groove is cut
along the periphery of the bearing block to fit a rubber O-ring type seal. This seal
has an inner diameter of 58 mm and a thickness of 3 mm.
8. Angular position plate: Using a reflective sensor (the measuring, data acquisition
and control system will be discussed thoroughly in the next section) the angular
position of the rotor is determined. A circular plate with four notches equispaced
along the plate’s periphery is secured to the shaft coupler and therefore rotates
inside the generator housing at the same rotational speed as the rotor. The
reflective sensor is installed immediately behind the plate, and hence, on each
rotation, the signal from the sensor changes eight times. By logging the amount of
signal perturbations per unit time, the rotational speed of the rotor can be
computed.
9. Brushless DC motor enclosure: A brushless DC motor (Warhead 3545-1850KV
EDF) is secured with four 3 mm countersunk bolts within the aluminum enclosure.
Configured as a generator, this motor serves to provide resistive torque by
applying an electrical load in order to control the turbine’s rotational speed. The
selection of this motor was based on its local availability, its high power to size
ratio and the expected operating conditions during wind tunnel experiments
(which were determined in the previous section using BEM theory). In fact, it is
suspected that this motor may not provide enough resistive torque when testing
in the water flume and therefore limiting the range of experimental results to high
RPM values only. Thus, although perfectly suitable for wind tunnel experiments
(see Section 6), the projected water flume experiments probably require
replacement of said brushless DC motor with a stepper motor or small induction
motor of similar size designed for higher torque values at lower speeds.
10. Torque tube: A custom made torque sensor, which is, in principle, just a hollow
aluminum (stationary) shaft with four linear type strain gauges (or load cells) in a
full bridge configuration oriented at 45 degrees, equispaced along the shaft’s
circumference. The angular displacement of the shaft due to applied torque causes
elongation (and compression) of the strain gauges glued onto its surface, thereby
creating an unbalance in the bridge which is expressed in millivolts. Said

16RET1012
43
unbalance can be related to the magnitude of the torque present by calibrating the
setup (discussed in the next section). Note that this torque sensor is rigidly
connected to both the rear lid of the turbine housing and the motor enclosure and
therefore does not rotate with the turbine’s rotor; torque is transmitted via
electromagnetic forces between the stator and the rotor within the motor. The outer
diameter of the shaft equals 12.4 mm, while its wall thickness is only 0.2 mm. These
dimensions are based on the expected operating conditions, maximizing angular
displacement without plastic deformation at maximum torque. Details on the
calibration of the torque sensor are provided in the next section.
11. Torque sensor support/rear lid: The rear lid of the turbine housing also serves as
a fixed support for the torque sensor. It is manufactured from aluminum, such that
it slides partially into the cylindrical part of the turbine housing and is secured in
place with hex bolts.
12. Rear O-ring seal: There is a small gap (= 0.05 mm) between the concentric surfaces
of the rear lid and the turbine housing, potentially exposing the generator and
measuring devices inside to water contamination. To prevent water from leaking
into the turbine housing, a square groove is cut along the periphery of the rear lid
to fit a rubber O-ring type seal. This seal has an inner diameter of 64 mm and a
thickness of 3 mm.
13. Turbine housing: The cylindrical part of the turbine housing is made from
aluminum, has an outside diameter of 75 mm and a length of 219.50 mm. The
diameter has been minimized as much as possible, in order to reduce any
interference with the flow through the rotor blades.
14. Turbine support: An airfoil-shaped support structure. The turbine housing is fixed
onto the support by means of hex bolts. To ensure leak-free operation a bead of
silicone should be applied between the mating surfaces of respective parts. Electric
wiring passes through the hollow support structure via two holes into the turbine
housing. It should be noted that this component was not manufactured as part of
this project, but was already present in the one of the laboratories at the University
of Brasilia.

In order to provide an improved understanding of the modified design of the lab-scale


turbine, supplementary 3-dimensional photorealistic renderings are included hereinafter,
see Figure 5.2.

All components of the lab-scale turbine, except the rotor and the flow deflector, were
manufactured via several machining operations in an automotive laboratory of the
University of Brasilia in Gama within a timeframe of two weeks. The rotor and flow
deflector, on the other hand, have been printed at a 3D fabrication laboratory on the main
campus of the same University. Although the manufacturing process of the lab-scale
turbine is not discussed in detail herein, it is deemed necessary (and informative) to make
the following comments on the printing and machining operations involved:

i. The rotor and flow deflector were printed with ABS filament, using a da Vinci 1.0
PRO 3D printer. The layer resolution was set at 20 microns, which is the highest
setting. Considering maximum structural strength and laboratory experiments on
the dynamic behavior of the turbine (this was discussed in Section 1.2), the rotor

16RET1012
44
(a)

(b)
Figure 5.2 Photorealistic renderings of the developed lab-scale HK-10 hydrokinetic turbine.

was printed completely solid, thereby ensuring a homogeneous mass distribution.


Note that due to the overhanging rotor blades, it was necessary to print a
supporting structure to allow for successful printing, as can be seen in Figure 5.3a.
The flow deflector, on the other hand, is a stationary component which mass does
not affect the dynamic behavior of the turbine, it is not considered a structural part
and has a relatively simple shape. Therefore, the fill density was set at 20% using
a honeycomb fill pattern and no supporting structures were required. The printing
time for the rotor was approximately 27 hours, whereas the flow deflector was
printed in about 7 hours.
ii. Although printing with 20 micron layer thickness is considered high quality, there
is still a noticeable roughness to the components’ surfaces. Moreover, when using
non-soluble support material (which was the case), additional surface
imperfections are introduced which are not desired. In order to create a smooth
surface, post processing of the 3D prints was done; a polyester resin based filler
material was applied to the surfaces, whereafter a series of sanding steps,
progressively using finer grits, were conducted, followed by priming and painting
of the components, see Figures 5.3b and 5.4a.
iii. Most machining operations have been performed with a manual metal lathe and
milling machine, using a number of different accessories and cutting tools. Raw
materials (aluminum, brass and steel) were locally available in Brasilia. The
machined parts are shown in Figure 5.4b.

16RET1012
45
(a) (b)
Figure 5.3 (a) Printed rotor with support material, (b) Filler material applied to the rotor and stator,
whereafter sanding operations were performed to smooth out the surfaces.

(a)

(b)
Figure 5.4 (a) Painted 3D printed parts. (b) Machined (and partially assembled) parts.

16RET1012
46
5.2 Data acquisition and control system
5.2.1 Measuring system
Within the context of this project there were principally only two important variables to
be measured with respect to the turbine’s operation: shaft torque and rotational speed. As
stated earlier, these measurements were made using a reflective optical sensor and strain
gauges, respectively. Figure 5.5a shows a photo of the optical sensor (Vishay® TCRT5000)
installed within the turbine housing. This sensor includes an infrared emitter (sender) and
phototransistor (receiver) and has a peak operating distance of 2.5 mm; the sensor was
positioned accordingly. With the angular position plate turning, the signal from the optical
sensor get interrupted four times within one full rotation. Via a fairly simple electrical
circuit (see Figure 5.5b) and the use of an Arduino MEGA 2560 microcontroller, the
number of interruptions per unit time are registered, from which the semi-instantaneous
rotational speed of the turbine can be computed.

TCRT5000

(a)

b)
Figure 5.5 (a) Optical sensor installed within turbine housing. (b) The electrical diagram constructed to employ
the TCRT5000 optical sensor for RPM measurements.

Shaft torque is measured by means of linear type strain gauges (90 Ω), glued onto the
torque tube with a cyanoacrylate based adhesive (see Figure 5.6a), connected in a full
Wheatstone bridge configuration. Their orientation is such that an angular displacement
of the torque tube due to turbine shaft torque causes two of these strain gauges to stretch,
whereas the other two become compressed, thereby creating an unbalance in the bridge
which can be measured in millivolts. This voltage signal is amplified via a 24-bit precision

16RET1012
47
(a)

(b)
Figure 5.6 (a) Strain gauges glued onto the torque tube. (b) Full Wheatstone bridge configuration and signal
amplification using an ADC.

analog-to-digital converter (HX711 ADC), which enables the Arduino MEGA 2560
microcontroller to detect changes in the strain gauge resistances, see Figure 5.6b.

The torque sensor was calibrated using a setup as shown in Figure 5.7, where a known
mass (𝑚 = 50.496 g) was suspended at different positions along a lever arm fixed on the
generator enclosure. As a result, torques of known magnitude are induced in the torque
tube equipped with the strain gauges. By relating the different torque values to their
corresponding voltage outputs from the ADC module, a scaling parameter can be
calculated which is used to translate between voltage signals and torque values. Following
said procedure, a plot was constructed which includes the measured torque values and
the calculated torque values as well for comparison (see Figure 5.8). Clearly, there exists a
linear relationship between the measured torque and the distance from the pivoting point,
as was confirmed with a regression analysis (𝑅 2 ≅ 1). Moreover, with deviations less
than 0.7 Nmm, there is an excellent agreement with the calculated values. Considering the
expected torque magnitude during experiments, it was concluded that the custom made

16RET1012
48
Figure 5.7 Calibrating the torque sensor with a know weight at different positions along a lever arm fixed to
the generator housing.

150

100

50
τ [Nmm]

-50

-100

-150
-250 -200 -150 -100 -50 0 50 100 150 200 250
d [mm]

Calculated data Experimental data

Figure 5.8 Measured and calculated torque values related to a mass suspended at different positions from the
pivoting point in the calibration setup.

torque sensor provides sufficient accuracy and is therefore suitable for its intended
purpose.

5.2.2 RPM Control system


In order to experimentally determine the power curve of the lab-scale HK-10 turbine, it is
necessary to control its rotational speed around a number of arbitrary set points within
the range of interest. In principle, such control can be done by modulating the time during
which a resistive load, which induces so-called braking torque, is connected to the
generator. The latter was made physically possible by employing a MOSFET transistor
(IRFZ48N) as a switching device, with its on-time regulated using a pulse-width
modulation (PWM) technique, in the electrical circuit as shown in Figure 5.9. The PWM
method was implemented using an Arduino MEGA 2560 at a switching frequency
of 490 Hz, whereas the pulse width (or pulse duration) was continuously adjusted in
accordance with the error existing between the actual RPM value and its set point via a

16RET1012
49
proportional-derivative-integral (PID) control scheme. Note that the PID control
parameters have been optimized/determined using a trial-and-error approach. It should
be noted, however, that these control parameters are specific to certain operating
conditions and therefore are not necessarily identical for the wind tunnel and water flume
experiments. The continuously variable electrical load system just described was
independently implemented for two line pairs of the brushless DC generator, with relays
(G5V-2 5DC) connected parallel to the transistors in the respective circuits. The relays are
energized, replacing the transistors’ function, whenever the PWM control is operating at
100% duty cycle. Basically, there are three possible braking modes: i) PWM control on one
line pair and zero load on the other line pair, ii) relay energized on one line pair and PWM
control on the other line pair, and iii) both relays energized, thereby providing maximum
braking torque (at which point there is effectively no control anymore).

Figure 5.9 RPM control system diagram with the transistor and relay connected in parallel, depicted for one
line pair of the brushless DC generator.

It should be noted that this RPM control system was largely based on the existing control
system developed by the Brazilian researchers (Brasil & Lavaquial, 2017) for their research
on different horizontal axis lift based rotors in their wind tunnel setup in Brasilia. Some
adjustment have been made to the original Arduino sketch, in which the PID signal got
tweaked in order to successfully control the turbine’s rotational speed; instead, the PID
control parameters have been optimized, as was previously stated, which resulted in a
relatively faster and more accurate control system with respect to its predecessor.

16RET1012
50
6 EXPERIMENTAL STUDY
6.1 Wind tunnel experiments with hydrokinetic turbines
The lab-scale model developed as part of this project is primarily intended for actual
underwater experiments, as opposed to the previously existing model, which is only
suitable for wind tunnel experiments. Nevertheless, for validation purposes and in order
to create a baseline for the projected experiments in water streams, it is found necessary
to firstly test the developed lab-scale model in a wind tunnel. By comparing the wind
tunnel results with future water flume results, the feasibility of using a wind tunnel for
experimental evaluation of horizontal axis lift based hydrokinetic turbines through the
application of dimensional analysis and scaling laws, such as will be presented in the next
paragraph, can be assessed better. Thus far, said feasibility has only been assessed by
comparing (transposed) experimental results with numerical results from CFD and BEM
simulations under steady state conditions. Although these computational methods have
proven to be very powerful tools for predicting flow characteristics and turbine
performance, respectively, they are not a substitute for actual physical experiments. In
fact, these methods do have limitations and do not always correctly predict (with an
acceptable level of uncertainty) real behavior, as was, for example, pointed out in Section
4.2, regarding the tip-speed-ratio at which the power coefficient of the HK-10 turbine is
maximum according the BEM theory.

6.1.1 Brief description of the wind tunnel testing facility, UnB


The University of Brasilia (UnB) has an open-loop closed-section wind tunnel (see Figure
6.1) installed in its Energy and Environment laboratory in which the developed lab-scale
turbine has been tested. Air flow is produced by means of a fan located at the exit of the
tunnel, effectively sucking air into the tunnel through a honeycomb mesh at the inlet. The
fan is driven by a 10 kW induction motor, with a variable-frequency drive (VFD) to control
its rotational speed. The wind speed can be varied between 0 m⁄s and approximately
20 m⁄s via a feedback control system using information from a differential pressure
sensor connected to a pitot tube inside the test section. The turbulence intensity level in
the tunnel has been measured using a hot-wire anemometer and turns out to be less than
1% within the range possible wind speeds (Brasil & Lavaquial, 2017). The wind tunnel has
total length of 15 m, whereas its test section dimensions are 1.2(W) × 1.2(H) × 2(L) m.

Figure 6.1 Open-loop wind tunnel in the Energy and Environment laboratory on the main campus of the
University of Brasilia, UnB.

16RET1012
51
6.1.2 Interpretation of wind tunnel results using dimensional analysis
The geometry of the HK-10 hydrokinetic turbine was optimized for specific operating
conditions and its size was based on a desired power output; information which has been
provided in Section 2.1. Particularly because of its relatively large rotor (𝐷 = 2.1 m), it is
practically not possible to use the prototype model to perform controlled experiments in
either the wind tunnel or the water flume, given the dimensions of their respective test
sections. This situation demands for a (smaller) scale model, such as developed as part of
this project. Moreover, the testing conditions in a wind tunnel being inherently different
from the design conditions imposes additional challenges.

Although scale models allow for testing of a design, it is important to determine the correct
testing conditions such that the experimental results are applicable to the real design, i.e.,
criteria for similitude must be met. The concept of similitude (or similarity) within the
field of fluid mechanics, states that a (scale) model is similar to the real application if they
share geometric, kinematic and dynamic similarity. While the developed lab-scale model
has exactly the same shape as the prototype (it is only downscaled with a factor of 13.125)
and thus is geometrically similar, conditions for kinematic and dynamic similarity are not
attainable 5 in the wind tunnel because of the large deviations in Reynolds number
between the laboratory conditions and the design conditions. Furthermore, it must be
pointed out that due to the complexity of flow phenomena dependent on the Reynolds
number, such as turbulent-lamina transitions and boundary layer detachment, findings
from studies on the relative differences between experimental results using lab-scale
models and prototype models, respectively, cannot be (easily) generalized.

A semi-empirical model, proposed in Ref. (Brasil & Lavaquial, 2017), provides a means for
transposing wind tunnel experimental results with the lab-scale model, to “prototype
results”. More specifically, in this model, the power coefficient of a lab-scale hydrokinetic
turbine is related to that of the prototype, by performing a dimensional analysis involving
BEM simulation results. Because it is of crucial importance to the conclusions that can be
drawn from the obtained results, the derivation of said model will be discussed next.

It can be argued that the power, P, produced by a hydrokinetic turbine is dependent on


its characteristic diameter, 𝐷, its rotational velocity, 𝜔, and the fluid density, 𝜌, viscosity,
𝜇 and velocity of the flow, 𝑢∞ , in which it is submerged, i.e.,

𝑃 = 𝑓(𝐷, 𝜔, 𝜌, 𝜇, 𝑢∞ ) (6.1)
Application of the Buckingham П theorem in order to transform the functional
relationship provided by equation (6.1) into a dimensionless form yields the following П
groups (note that 𝐷, 𝜌, and 𝑢∞ are selected as repeating variables),
𝑃 𝜇 𝐷𝜔
Π1 = 3 Π2 = Π3 = (6.2)
𝜌𝐷 2 𝑢∞ 𝜌𝐷𝑢∞ 𝑢∞

5Kinematic similarity is actually attainable, but only at speeds far above the maximum wind speed
of the wind tunnel available in Brasilia (see previous section); it can be easily shown that the speed
inside the tunnel should be around 517 m/s, which is above the thermodynamic speed of sound in
air at atmospheric conditions and therefore compressible flow phenomena may potentially occur,
complicating or even impede the analysis. 52

16RET1012
Interestingly, the first dimensionless group, Π1 equals the definition for the power
coefficient, 𝐶𝑃 (see equation (2.1)), Π2 is the reciprocal of the Reynolds number, Re, and Π3
corresponds to the tip-speed-ratio, TSR. Consequently, the dimensionless relationship
between the pertinent variables can be written as

𝐶𝑃 = 𝑓(Re, TSR) (6.3)


By performing the transposition analysis between the prototype and the lab-scale model
only at the nominal operating condition of the turbine (corresponding to 𝛼 = 6.5°, refer to
refer to Figure 4.4), it can be stated that TSRprot = TSRmod , and therefore,
𝐶𝑃,prot Reprot
= 𝑓( ) (6.4)
𝐶𝑃,mod Remod
3
Dividing equation (2.27) by 1⁄2 𝜌𝑢∞ 𝜋𝑅 2, which is the kinetic energy contained in the flow
through the (imaginary) actuator disc, yields another expression for the differential power
coefficient,

𝑑𝐶𝑃 = 8TSR2 (1 − 𝑎)𝑎′ 𝑟 ∗ 𝑑𝑟 ∗ (6.5)


where 𝑟 ∗ ≡ 𝑟⁄𝑅. Note that angular and axial induction factors, respectively 𝑎 and 𝑎′ , can
be computed using BEM theory. Since TSRprot = TSRmod, the ratio between the differential
power coefficients of the prototype and model, respectively, can be written as

𝑑𝐶𝑃,prot (1 − 𝑎)𝑎′ 𝑟 ∗ 𝑑𝑟 ∗ |prot


= (6.6)
𝑑𝐶𝑃,mod (1 − 𝑎)𝑎′ 𝑟 ∗ 𝑑𝑟 ∗ |mod
In essence, this ratio can be computed for each blade element of the rotor, nevertheless,
the model proposed in Ref. (Brasil & Lavaquial, 2017) disregards the radial dependency
of the induction factors and assumes that

𝐶𝑃,prot 𝑑𝐶𝑃,prot
= | (6.7)
𝐶𝑃,mod 𝑑𝐶𝑃,mod 𝑟=1⁄
2𝑅

The differential power coefficients, 𝑑𝐶𝑃,prot and 𝑑𝐶𝑃,mod , in the right-hand side of equation
(6.7) have been computed using the BEM method for the prototype turbine and the
developed lab-scale turbine, respectively, as can be seen in Figure 6.2. At the nominal
operating condition, corresponding to TSR = 1.61 ∧ 𝛼 = 6.5°, the value of 𝐶𝑃,prot ⁄𝐶𝑃,mod
was computed to be 1.63. Based on this value, a semi-empirical law has been proposed in
Ref. (Brasil & Lavaquial, 2017), which allows for transposition of wind tunnel
experimental data at any Reynolds number and tip-speed-ratio (recall that the validity of
the computed ratio is actually limited to the nominal operating condition only):

𝐶𝑃,prot Reprot 𝑚
=( ) (6.8)
𝐶𝑃,mod Remod
where 𝑚 is a scaling factor computed as follows
𝐶𝑃,prot
log [𝐶 ]
𝑃,mod
𝑚= (6.9)
𝑅𝑒prot
log [
𝑅𝑒mod ]

16RET1012
53
1.8 16

14
1.7
𝑑𝐶𝑃,prot
= 1.63
𝑑𝐶𝑃,mod 12
dCPprot/dCPmod [-]

1.6
10

α [˚]
1.5 8
𝛼 = 𝛼𝑜𝑝𝑡 = 6.5°
6
1.4
TSR nom = 1.61

1.3
2

1.2 0
1.2 1.4 1.6 1.8 2.0 2.2 2.4 2.6
TSR [-]

dCPprot/dCPmod Angle of attack, α

Figure 6.2 The ratio between 𝑑𝐶𝑃, prot and 𝑑𝐶𝑃, mod has been computed for different values of TSR in order
the perform a transposition between the lab-scale model and the prototype. At the nominal operating
condition (TSR = 1.61) said ratio has a value of 1.63.

With Reprot = 1.1 ∙ 106 and Remod = 3.0 ∙ 104 (see the next section for details on the
experimental conditions) and 𝐶𝑃,prot ⁄𝐶𝑃,mod = 1.63, the scaling factor, 𝑚, takes on a value
of approximately 0.136. Note that the Reynolds numbers have been computed with the
characteristic linear dimension set equal to the chord length at 𝑟 ∗ = 0.5, for the respective
rotors.

Using equation (6.8) results obtained from the wind tunnel experiments have been
transposed to prototype results, in an attempt to account for the dynamic dissimilarity
between the lab-scale setup and the prototype. It should however be pointed out that
application of this method is considered somewhat unreliable, as it rests on some weakly
founded assumptions. Frankly, in order to obtain good experimental results on the
hydrodynamic performance of the turbine, dynamic similarity must be guaranteed, which
is practically only possible when testing at controlled conditions in a water flume such as
proposed herein. Nevertheless, the results from the transposition analysis are presented
in the next section, to serve as a basis for comparison between the discussed method and
actual water flume experiments with the developed lab-scale HK-10 turbine in the future.

16RET1012
54
6.2 Experimental procedure, results and discussion
The developed lab-scale HK-10 turbine has been experimentally tested in a wind tunnel
installed in the Energy and Environment laboratory at the University of Brasilia (refer to
Section 6.1). The experimental procedure can be summarized as follows:

i. The airfoil-shaped turbine support (part number 14) was bolted onto the bottom
of the wind tunnel’s test section (see Figure 6.3), with all wiring guided through a
hole up to the data acquisition and control unit, which was fixed on an
instrumentation panel on the outside of the tunnel.
ii. Electrical connections were made according to the diagrams shown in Section 5.2,
i.e., the leads from strain gauges were connected in a full-bridge configuration to
the HX711 module, the three wires from the brushless DC motor where connected
to the PWM control system, as was the TCRT5000 optical sensor.
iii. The data acquisition and control system was powered and the Arduino MEGA
2560 microcontroller was connected to a computer via a serial communication port
in order read live data (through a serial monitor) and command RPM set points.
iv. The wind tunnel’s fan was started via a SCADA system, in which a wind speed of
15 m⁄s was demanded. Note that due to extra driveline components (intermediate
shaft, ball bearings and a coupling), there are additional mechanical losses and
more mass inertia, which requires increased starting power. In fact, at wind speeds
below 13 m/s, the turbine would not start rotating by itself (even with no electrical
braking applied). Mechanical losses have been quantified, as will be discussed
later.
v. As soon as a stable wind velocity was reached inside the tunnel, the torque and the
rotational speed (approximately 3600 revolutions per minute) were recorded from
the serial monitor. Next, the rotational speed of the turbine was gradually reduced
by 200 revolutions per minute each time (by changing the RPM control set point
via the serial communication port) until no further braking could be applied, at
which point the rotational speed was at approximately 800 revolutions per minute.
For each rotational speed the corresponding torque value was recorded. It should
be noted that the respective RPM and torque values were actually recorded for a
period of approximately 30 seconds, only after the deviation between the
measured RPM value and the set point value reached values below 50 revolutions
per minute; due to oscillations about the set point value as a consequence of the
PID controller action and mechanical vibrations, there is a continuous variation in
the magnitude of the measured quantities. Since both the torque and RPM
measurements during the said 30 second period are assumed to be normally
distributed, the medians of the respective datasets were employed to compute the
power curve of the lab-scale HK-10 turbine.
vi. After all measurements had been done, the wind tunnel fan was turned off and the
data acquisition and control system was shut down.

16RET1012
55
Figure 6.3 The lab-scale HK-10 turbine set up in the wind tunnel at the University of Brasilia.

As was stated before, there are mechanical energy losses due to bearing friction. These
losses have been experimentally quantified by running the generator as a motor without
the rotor installed, therefore being able to measure the friction torque using the same data
acquisition system as was presented in Section 5.2. In order to run the brushless DC motor,
a 30 A Electronic Speed Control (ESC) driver was employed together with a rotary
potentiometer (1 kΩ − 500 kΩ) to adjust the rotational speed. In principle, the frictional
torque should at least be measured between 800 and 3600 RPM (refer to the experimental
procedure), however, due to the motor’s high KV rating (1850) and its voltage cut-off limit
( 11.1 V ), it was not possible to perform measurements below 2300 RPM. From the
experimental results shown in Figure 6.4, it can be seen that there seems to exists a linear
relationship between the friction torque and the turbine’s rotational speed on the
measured domain. It is assumed that this relationship ( 𝜏fric = 0.001 ∙ 𝑛 + 9.113 ; this
equation corresponds to the trendline in Figure 6.4) is valid down to 800 RPM and
therefore the results have been extrapolated to make up for the lack of experimental data
points. Clearly, the mechanical energy losses due to bearing friction are quite significant
and therefore must be accounted for when computing the power curve for the HK-10 lab-
scale model. The power coefficient, 𝐶𝑃 and the tip-speed-ratio, TSR, have been computed
from the experimental results using equations (2.1) and (2.2), respectively, see Figure 6.5.

16RET1012
56
25

20
τfric [Nmm]

15

10

0
2000 2500 3000 3500 4000 4500 5000 5500 6000 6500 7000
RPM [rot/min]

Experimental data Linear (Experimental data)

Figure 6.4 Measured friction torque for RPM values ranging from 2300 up to 6500.

0.25

0.2

0.15
CP [-]

TSR = 1.61

0.1

0.05

0
0 0.5 1 1.5 2 2.5
TSR [-]

Power Coefficient, CP (exp. data) Power Coefficient, CP (corrected data)

Figure 6.5 Experimentally determined power curve for the lab-scale HK-10 turbine, without the correction
applied for mechanical energy losses, and the corrected data. Clearly, friction losses in the lab-scale model are
significant.

16RET1012
57
Figure 6.5 shows the experimentally determined power curve for the lab-scale HK-10
model inside the wind tunnel (see the data with the round markers), without accounting
for mechanical energy losses. From a qualitative point of view it can be stated that the
curve behaves such as expected (e.g., when comparing to the numerical results), but from
a quantitative perspective the results are noticably different. These differences are largely
eliminated when the experimental data points are corrected to account for the mechanical
energy losses. The corrected data points are included in Figure 6.5 as well (see the data
with the triangular markers); the top of the power curve is close to 𝐶𝑃 = 0.25 such as
predicted by the BEM model and its location is actually at approximately TSR = 1.61,
which coincides with the design condition (note that the latter observation is different
from what the BEM model predicts, which will be discussed shortly).

The windtunnel results have been transposed to prototype results using equation (6.8)
with the scaling parameter, m, set to 0.136 (refer to Section 6.1.2). The results of this
exercise are included in Figure 6.6 (see the data points with square markers), along with
the previously presented BEM simulation result for comparison. Both the experimental
results and the BEM results indicate that the maximum power coefficient for the HK-10
turbine is approximately 0.38, which only marginally differs from the design condition
(𝐶𝑃 = 0.4). Moreover, there is a fairly good agreement between experimental results and
BEM results at relatively low tip-speed ratios, which however deteriorates quickly starting
from TSR = 1.2 . In fact, the BEM model predicts the top of the power curve at a
significantly larger tip-speed ratio, as was previously discussed at the end of Section 4; the
reader is referred to said discussion with regard to this observation.

0.45

0.40

0.35

0.30

0.25
CP [-]

TSR = 1.61

0.20

0.15

0.10

0.05

0.00
0.0 0.5 1.0 1.5 2.0 2.5 3.0
TSR [-]

Power Coefficient, CP (Prototype, Transposed experimental values)

Power Coefficient, CP (Prototype, BEM)

Figure 6.6 BEM simulation results versus the transposed (and corrected) experimental values from the wind
tunnel experiment.

16RET1012
58
7 CONCLUSIONS AND RECOMMENDATIONS
An experimental investigation of flow devices such as hydrokinetic turbines by means of
lab-scale models always has to be accompanied by an analysis which determines the
required testing conditions to guarantee dynamic similitude. At the University of Brasilia
hydrokinetic turbine characteristics are experimentally investigated in a wind tunnel.
However, due to the extremely large deviation existing between the Reynolds numbers of
the flow inside the wind tunnel and the actual flow through the turbines, respectively, it
is practically impossible to achieve dynamic similitude. Consequently, results from the
wind tunnel experiments with the lab-scale HK-10 turbine are principally not applicable
to the real design, despite the fact that conditions for geometric similarity are met.
Nevertheless, in an attempt to account for said dynamic dissimilarity, a dimension
analysis relating BEM simulation results of the lab-scale turbine to that of the prototype
turbine, respectively, was conducted, resulting in a semi-empirical scaling law which has
been employed by researchers at the University of Brasilia to transpose the wind tunnel
results. It was however concluded that the used method does not correctly account for the
dynamic dissimilarity. In fact, i) Reynolds number dependent complex flow phenomena
are excluded from the analysis, although these may have a considerable effect on the
experimental results; frankly, these complexities cannot be easily investigated from a
theoretical point of view, ii) the developed scaling law deceptively assumes a homogenous
distribution of axial and angular induction factors along the blade length; actually there is
a significant variance, as became evident from the BEM simulation results and therefore
scaling factors should have been computed for each blade element separately, iii) the
transposition analysis is completely dependent on the BEM method, which however
clearly behaves erratically from a quantitative perspective, thereby rendering the
transposed results rather unreliable.

In order to guarantee dynamic similitude and additionally provide a means for studying
the transient behavior of hydrokinetic turbines, a water flume testing setup is proposed.
Conveniently, the Anton de Kom University of Suriname already has a water flume testing
facility, but it however does not meet the technical requirements to test hydrokinetic
turbines as was discussed in Section 3.2. Specifically, i) the pumping capacity of the testing
facility is insufficient to reach a practical water level in the flume at desired flow speeds,
ii) water is fed into the flume in a very inefficient manner, resulting in a unfavorable test
condition with respect to the turbulence level, moreover, kinetic energy is lost due to high
velocity gradients, particularly at the inlet of the flume.

A rather simple reconfiguration/modification to the water feeding system allows for a


tangent propagation of the water flow into the flume, unarguably promoting a laminar
test condition. Additionally, flow straightening tubes just before the test section are
proposed in order to further reduce the magnitude of transverse velocities. The water level
is locally increased by means of hydraulic jump, induced by a flume width transition
from 1 m to 0.5 m, which is possible as long as the upstream flow is supercritical. Note that
such a flow constriction also supports a homogeneous velocity vector field. Application
of the energy conservation law has shown that it is possible to achieve flow heights
between approximately 30 and 40 cm, and flow speeds of ranging between 2 and 5.4 m/s,
in the test section, when throttling is applied in the water flume inlet pipes. In order to
provide more accurate results and to study the velocity vector field inside the test section

16RET1012
59
for the hydrokinetic turbines, a CFD simulation was conducted for a scenario with zero
throttling in the inlet pipes. The CFD simulation predicts the water level in the test section
at 28 cm, as opposed to 29.7 cm computed using the energy conservation method. The
flow speed, on the other hand, averages at approximately 4.6 m/s in the mid plane of the
test section according to the CFD results, whereas the energy conservation method
predicts a velocity of 5.4 m/s. This quite significant difference can be attributed to the fact
that the latter method does not consider kinetic energy losses. Moreover, it can be
concluded that the water level in the constricted zone of the flume is a much weaker
function of the specific energy, than the flow speed is, and therefore is less affected by
kinetic energy losses. With respect to the velocity vector field inside the test section, an
investigation of the CFD results has shown that transverse velocities are relatively small,
whereas for the main flow direction the velocity is homogenously distributed throughout
an area which is considered large enough for practical testing of hydrokinetic turbines,
when considering boundary layer influence. It should be noted however that there have
been no guidelines found in literature on blockage ratios for water flume experiments and
therefore a 10% blockage ratio, as is typical in wind tunnel experiments, has been adopted
for the purpose of sizing the lab-scale HK-10 turbine. The CFD simulation results are
qualitatively as expected, and even from a quantitative point of view seem to be
reasonable, when comparing to the energy conservation method (as far as can be
compared), however, it must be pointed out that the results are still questionable as no
grid dependency study could be conducted. Moreover, the exact influence of the
adaptations on the test conditions is not studied/quantified and therefore no solid
statements can be made on this matter.

BEM computations have been performed for both the prototype turbine at design
conditions and the lab-scale HK-10 turbine at wind tunnel testing conditions. The
computed power curves are qualitatively as expected with maximum power coefficients
of 0.38 and 0.25, respectively. However, these maxima are located at tip-speed-ratios
different from the design conditions, which is quite interesting; experimental analysis of
the manufactured lab-scale turbine actually confirms the location of the maximum
approximately at a tip-speed-ratio of 1.61, which is the design value. It can therefore be
concluded that the actuator disk model correctly optimizes the blade geometry, but the
BEM model fails to make a proper prediction on its hydrodynamic behavior from a
quantitative point of view. It is suspected that the hydrodynamic coefficients for the
NACA 4415 profile are inaccurately computed by XFOIL at the simulated Reynolds
numbers. Unfortunately, this can only be confirmed with actual experiments on the lift
and drag coefficients, which is beyond the scope of this work.

The lab-scale HK-10 turbine was manufactured such that it allows for underwater testing
as intended, as opposed to its predecessor which due to its open generator housing is only
suitable for wind tunnel experiments. A fully enclosed aluminum generator housing with
seals and special bearings ensures leak-free operation. Moreover, a static flow deflector
promotes a fluent propagation of the flow through the rotor. It should be noted, however,
that the increased driveline components with respect to its predecessor significantly
increases mechanical energy losses, which should be considered when experimentally
investigating the performance of different rotors.

16RET1012
60
The obtained experimental results with the lab-scale HK-10 turbine inside a wind tunnel
have been corrected for friction torque and thereafter have been transposed using the
semi-empirical scaling law as mentioned previously. From analyzing these transformed
results it is seen that a maximum power coefficient of about 0.38 (at a tip-speed-ratio of
1.61) can be achieved with the HK-10 turbine, as was also predicted by the BEM model.
However, for reasons discussed earlier, these semi-experimental results should be used
with much prudence and in principle merely serve as a baseline for comparison in
following studies, within the context of this work. By all means, the results cannot be used
to validate the BEM model, as was wrongfully done in a previous work; after all, the
results are partially dependent on the BEM model itself.

Based on the outcomes from this study it is recommended to physically implement the
proposed modifications to the water flume testing facility at the Anton de Kom University
of Suriname on short notice, in order to enable experiments with hydrokinetic turbines in
water streams to better understand their hydrodynamic behavior. It should be noted,
however, that although the proposed modified water inlet and the flow straightening
tubes inarguably positively affect the flow conditions in the test section, their exact
influence must still be studied, ideally before physical implementation, to allow for
improvements/optimization.

Furthermore, it is recommended to experimentally determine the lift and drag coefficients


as a function of the angle of attack for the NACA 4415 profile, both in a wind tunnel and
a water flume, following the hypothesis on the erratic behavior of the BEM model as
discussed. Also, the validity of Buhl’s empirical relation for axial induction factors larger
than 0.4 (at which point momentum theory breaks down) should be investigated for
hydrokinetic turbines; it should be kept in mind that the relation was primarily intended
for applications with wind turbines.

16RET1012
61
16RET1012
62
REFERENCES
Alanne, K. & Saari, A., 2006. Distibuted energy generation and sustainable development.
Renewable and sustainable energy reviews, 10(6), pp. 539-558.

Anon., sd NACA 4415 airfoil. Urbana(Illinois): UIUC Applied Aerodynamics Group,


Department of Aerospace Engineering, University of Illinois.

Anyi, M. & Kirke, B., 2010. Evaluation of small axial flow hydrokinetic turbines for remote
communities. Energy for sustainable development, 14(2), pp. 110-116.

Bagher, A. M., Vahid, M., Mohsen, M. & Parvin, D., 2015. Hydroelectric Energy
Advantages and Disadvantages. American Journal of Energy Science, 2(2), p. 17.

Benini, E., 2004. Significance of blade element theory in performance prediction of marine
propellers. Ocean Engineering, 31(8-9), pp. 957-974.

Brasil, A. J. et al., 2017. Hydrokinetic propeller turbines: How many blades?. American
Journal of Hydropower.

Brasil, A. & Lavaquial, J., 2017. Projeto Hydro-K: Systemas de Turbinas hydrocinéticas em
arranjo flutuante e modular, Brasília: Universisade de Brasília.

Chapman, J. C., Masters, I., Togneri, M. & Orme, J. A., 2013. The Buhl correction factor
applied to high induction conditions for tidal stream turbines. Renewable Energy, Volume
60, pp. 472-480.

Chaudhry, M. H., 2007. Open-channel flow. sl:Springer Science & Business Media.

Chen, S., Chen, B. & Fath, B. D., 2015. Assessing the cumulative environmental impact of
hydropower construction on river system based on energy network model. Renewable and
Sustainable Energy Reviews, Issue 42, pp. 78-92.

Chen, T. Y. & Liou, L. R., 2011. Blockage corrections in wind tunnel test of small horizontal
axis wind turbines. Experimental Thermal and Fluid Sciences, 35(3), pp. 565-569.

da Silva, H. P., Blanco, C. J., Mesquita, A. L. & Brasil, A. J., 2017. Assessment of
hydrokinetic energy resources downstream of hydropower plants. Renewable Energy,
Volume 101, pp. 1203-1214.

DelSontro, T. et al., 2010. Extreme methane emissions from a Swiss hydropower reservior:
contribution from bubbling sediments. Environmental Science and Technology, 44(7), pp.
2419-2425.

Els, R. H. & Brasil, A. C., 2015. The Brazilian experience with hydrokinetic turbines. Energy
Procedia, Volume 75, pp. 259-264.

Els, R. H., Miranda, A. R., Brasil, A. J. & Echeverry, S. M., 2018. Hydrokinetic Energy
Conversion - State of the Art and Perspectives in Brazil. Rio de Janeiro, SDEWES.

Froude, W., 1878. On the elementary relation between pitch, slip and propulsive. Trans.
Roy. Inst. Naval Arch., 19(47), pp. 45-57.

16RET1012
63
Giles, J., 2006. Methane quashes green credentials of hydropower. sl:sn

Glauert, H., 1983. The elements of aerofoil and airscrew theory. sl:Cambridge University Press.

Goldthau, A. & Sovacool, B. K., 2012. The uniqueness of the energy security, justice and
governance problem. Energy Policy, Volume 41, pp. 232-240.

Guney, M. S. & Kaygusuz, K., 2010. Hydrokinetic energy conversion systems: A


technology status review. Renewable and Sustainable Energy Reviews, 14(9), pp. 2996-3004.

Hansen, M. O. L., 2008. Aerodynamics of Wind Turbines. 2nd red. London: Earthscan.

Khan, M. J., Bhuyan, G., Iqbal, M. T. & Quaicoe, J. E., 2009. Hydrokinetic energy
conversion systems and assessment of horizontal and vertical axis turbines and tidal
applications: A technology status review. Applied energy, 86(10), pp. 1823-1835.

Kolekar, N., Mukherji, S. S. & Banjeree, A., 2011. Numerical modeling and optimization of
hydrokinetic turbine. sl, ASME 2011 5th International Conference on Energy Sustainability,
pp. 1211-1218.

Kolekar, N., Vinod, A. & Banjeree, A., 2019. On Blockage Effects for a Tidal Turbine in Free
Surface Proximity. Energies, 12(17), p. 3325.

Lago, L. I., Ponta, F. L. & Chen, L., 2010. Advances and trends in hydrokinetic turbine
systems. Energy for sustainable development, 14(4), pp. 278-296.

Ledec, G. & Quintero, J. D., 2003. Good dams and bad dams: environmental criteria for site
selection of hydroelectric projects, sl: The World Bank.

Masters, I., Chapman, J. C., Willis, M. R. & Orme, J. A., 2011. A robust blade element
momentum theory model for tidal stream turbines including tip and hub loss corrections.
Journal of Marine Engineering & Technology, 10(1), pp. 25-35.

Mikkelsen, K., 2013. Effect of free stream turbulence on wind turbine performance (MSc thesis),
Trondheim: NTNU.

Mol, J., Merona, B. D. & Ouboter, P. E., 2007. The fish fauna of Brokopondo Reservior,
Suriname, during 40 years of impoundment. Neotropical Ichthyology, 5(3), pp. 351-368.

Rankine, W. J., 1865. On the mechanical principles of the action of propellers. Transactions
of the Institution of Naval Architects, Volume 6.

Reynolds, T. S., 1983. Stronger than a hundred men: a history of the vertical water wheel. sl:JHU
Press.

Shen, W. Z., Mikkelsen, R., Sorensen, J. N. & Bak, C., 2005. Tip loss corrections for wind
turbine computations. Wind Energy: An international journal for progress and applications in
wind power conversion technology, 8(4), pp. 457-475.

Smith, N., 1975. Man and water: a history of hydro-technology. sl:Scribner.

16RET1012
64
Sovacool, B. K., 2011. Evaluating energy security in the Asia pacific: Towards a more
comprehensive approach. Energy Policy, 39(11), pp. 7472-7479.

White, F. M., 2015. Fluid mechanics. sl:McGraw Hill.

16RET1012
65

You might also like