You are on page 1of 29

Edinburgh Research Explorer

Sensitivity analysis of a Venturi shaped structure for cross-flow


turbines

Citation for published version:


Gabl, R, Burchell, J, Hill, M & Ingram, D 2022, 'Sensitivity analysis of a Venturi shaped structure for cross-
flow turbines', Engineering Applications of Computational Fluid Mechanics, vol. 16, no. 1, pp. 2243-2269.
https://doi.org/10.1080/19942060.2022.2137850

Digital Object Identifier (DOI):


10.1080/19942060.2022.2137850

Link:
Link to publication record in Edinburgh Research Explorer

Document Version:
Publisher's PDF, also known as Version of record

Published In:
Engineering Applications of Computational Fluid Mechanics

General rights
Copyright for the publications made accessible via the Edinburgh Research Explorer is retained by the author(s)
and / or other copyright owners and it is a condition of accessing these publications that users recognise and
abide by the legal requirements associated with these rights.

Take down policy


The University of Edinburgh has made every reasonable effort to ensure that Edinburgh Research Explorer
content complies with UK legislation. If you believe that the public display of this file breaches copyright please
contact openaccess@ed.ac.uk providing details, and we will remove access to the work immediately and
investigate your claim.

Download date: 10. Aug. 2023


Engineering Applications of Computational Fluid
Mechanics

ISSN: (Print) (Online) Journal homepage: https://www.tandfonline.com/loi/tcfm20

Sensitivity analysis of a Venturi shaped structure


for cross-flow turbines

Roman Gabl, Joseph Burchell, Mark Hill & David M. Ingram

To cite this article: Roman Gabl, Joseph Burchell, Mark Hill & David M. Ingram (2022) Sensitivity
analysis of a Venturi shaped structure for cross-flow turbines, Engineering Applications of
Computational Fluid Mechanics, 16:1, 2243-2269, DOI: 10.1080/19942060.2022.2137850

To link to this article: https://doi.org/10.1080/19942060.2022.2137850

© 2022 The Author(s). Published by Informa


UK Limited, trading as Taylor & Francis
Group

Published online: 21 Nov 2022.

Submit your article to this journal

View related articles

View Crossmark data

Full Terms & Conditions of access and use can be found at


https://www.tandfonline.com/action/journalInformation?journalCode=tcfm20
ENGINEERING APPLICATIONS OF COMPUTATIONAL FLUID MECHANICS
2022, VOL. 16, NO. 1, 2243–2269
https://doi.org/10.1080/19942060.2022.2137850

Sensitivity analysis of a Venturi shaped structure for cross-flow turbines


Roman Gabl a , Joseph Burchella , Mark Hillb and David M. Ingram a

a Institute for Energy Systems, FloWave Ocean Energy Research Facility, School of Engineering, Uiversity of Edinburgh, Edinburgh, UK; b DHV
Turbines Ltd, Glasgow, UK

ABSTRACT ARTICLE HISTORY


Tidal energy is one of the world’s most predicable renewable energy sources and therefore holds Received 19 May 2022
great potential to be a valuable building block for the decarbonisation of electricity production. This Accepted 26 September 2022
paper focuses on a Venturi shaped duct structure (shroud) to accelerate the flow speed at a verti- KEYWORDS
cal axis tidal turbine utilising the low static pressure created at the exit of the shroud. This concept is Tidal turbine; vertical axis;
known as a Davidson Hill Venturi (DHV) turbine. By constructing the nozzle and diffusor using hydro- optimisation; Davidson Hill
foils, initial demonstrations indicate increased system efficiency. However, owing to the potential Venturi (DHV);
number of geometric and structural hydrofoil variations, only a general description of the location ANSYS -CFX ; SpaceClaim
of the hydrofoils is provided in order to facilitate modelling while allowing for future geometric vari-
ations to be devised. The conducted investigations focus on the influence of the nozzle and diffusor
sections as the main geometry variations, identifying the length component in the orthogonal direc-
tion as the dominant parameter. By modelling multiple combinations of these variables it is clear that
higher fluid velocities result in larger forces which must be supported by the devices structure. Small
adjustments to the reference geometries hydrofoil placement and spacing provided improvements
to the fluid flow. Thus, taking the slight alteration to the geometry as this papers main outcome, a
further in a 3D-simulation study, including turbine interaction and rotation, is to be completed to
fully characterise the systems benefits. The insights gained from this work will allow a reduction in
computational costs for the detailed optimisation and study into the adaption of the concept for a
wide range of (environmental) boundary conditions.

1. Introduction
In addition to the turbine structure itself, various
The kinetic energy of tidal currents is a renewable methods are available to increase the flow speed at the
energy source that is highly predicable and can provide turbine and therewith improve the overall performance
a key component for future fully decarbonised electrical of the turbine. For example, Z. Liu et al. (2019) inves-
energy production (Bahaj, 2013; Khojasteh et al., 2022). tigated four additional hydrofoils placed as a guide-
A wide range of turbine concepts are currently under vane diffuser using ANSYS -Fluent . H.-M. Wang
investigation (Roshanmanesh et al., 2020). For example, et al. (2020) used the software STAR-CCM+tm to investi-
they can be classified by the orientation of their rotation gate different ducts in a current flow limited by close side
axis in vertical and horizontal axis turbines (Behrouzi banks. This investigation was conducted as a 3D simula-
et al., 2016; Khan et al., 2009) or also into floating tion but without a turbine present. A three bladed vertical
and bottom mounted devices. A wide range of different axis turbine was simulated by El-Sawy et al. (2022) under
blades are used, including open center horizontal axis river conditions. An integration in a floating configura-
turbines (Belloni et al., 2017; Borg et al., 2020, 2021), tion is also possible (Hardisty, 2008). More commonly
twisted blades (Mosbahi et al., 2021) or deflector blades used for free stream turbines is a diffusor similar to those
(V. K. Patel & R. S. Patel, 2022). Blades can be either used for wind turbines (Arumugam et al., 2021; Noorol-
fixed, provide a variable pitch (Gu et al., 2018; H. Liu lahi et al., 2020; Nunes et al., 2020). These structures can
et al., 2020) and/or yaw angle (Modali et al., 2021; S.- be optimised for one single flow direction or allow for
q. Wang et al., 2018) or even change the shape (Pisetta bi-directional usage (Fleming & Willden, 2016), which
et al., 2022) to improve the efficiency further. More enables a fixed installation. Such ducted geometries can
classifications can be made using nacelle types, sup- be very simple structures but also include a variety of
porting structures and generators (Roshanmanesh et al., complex shapes and combined structures, as investigated
2020). by Huang et al. (2022) and Samadi et al. (2022). This

CONTACT R. Gabl roman.gabl@ed.ac.uk; J. Burchell j.burchell@ed.ac.uk; M. Hill mark@tidalenergy.com.au


© 2022 The Author(s). Published by Informa UK Limited, trading as Taylor & Francis Group
This is an Open Access article distributed under the terms of the Creative Commons Attribution-NonCommercial License (http://creativecommons.org/licenses/by-nc/4.0/), which permits
unrestricted non-commercial use, distribution, and reproduction in any medium, provided the original work is properly cited.
2244 R. GABL ET AL.

paper investigates a Venturi shaped structure that assem- ducted vertical turbine using a reduced 2D-approach and
bles multiple hydrofoils in order to accelerate the avail- simplified the turbine with an actuator disc approach
able flow through the vertical axis turbine (Kirke, 2011) in the numerical software OpenFOAM . In contrast to
instead of a single solid structure (Alidadi et al., 2008, this, the current paper focuses on an application of a
2011; Klaptocz et al., 2007). A summary of shapes and ducted vertical axis tidal turbine and introduces a basic
the influence of diffusors and horizontal axis turbines can description of the Venturi shaped shroud of the Davidson
be found in Nunes et al. (2020). Walker and Thies (2021) Hill Venturi (DHV) turbine. The supporting structure of
indicated that the percentage of failed tidal turbines was the turbine device is in part assembled using hydrofoils
higher, since the loads due to the increased velocity are arranged either side of the inlet, parallel to the axis of
higher. Nevertheless, they indicated that, with improved the turbine, improving the acceleration of the flow at the
materials and optimised concepts, the potential of ducted turbine. All numerical simulations are conducted with
turbines can be significant. the commercial software ANSYS -CFX . The geom-
Ongoing research challenges can be found in the etry generation is provided in the SpaceClaim spe-
continuing optimisation of blade geometry (Z. Wang cific IronPython as well as the open Python code (Gabl
et al., 2021; Zhu et al., 2020) as well as in the simulation et al., 2022). This ensures that the conducted variations
of the turbine itself, which can be simplified by various can be reproduced and expanded with any solver. The
approximations (Baratchi et al., 2020). Ke et al. (2020) methodology is described Section 2, which includes a ref-
compared a horizontal axis turbine with a range of diffu- erence geometry (Section 2.4) as well as the verification of
sors under a distributed inlet velocity as well as the inter- the 2D-numerical simulation (Section 2.6). A key output
action of multiple turbines in an array. Multiple vertical of the paper is the variation of the main geometry param-
axis turbine arrays are investigated by Sun et al. (2021). eters as well as the embedded hydrofoils (Section 3).
Ahmed et al. (2017) studied the influence of the velocity Section 4 provides a discussion of the chosen method-
distribution and compared it with field data and Badshah ological approach. The investigation is conducted with
et al. (2018) used a Fluid-Structure-Analysis to quantify reduced complexity of the numerical simulations and
the influence of a realistic velocity distribution in relation allows the identification of the optimum shape, which
to a homogeneous approach. will be numerically tested further in a fully 3D setup
The overall design of tidal turbines is strongly reliant to bring this concept closer to commercial deployment
on numerical simulations (Nachtane et al., 2020). Addi- producing fully predicable renewable energy from river,
tional, experimental investigations allow specific mea- canal and tidal flows.
surement data to be compared leading to improvements
within the numerical models and help quantify specific
influences, such as the arm effect for vertical axis tur- 2. Materials and methods
bines (Klaptocz et al., 2007; Li & Çalışal, 2010a, 2010b).
2.1. Overview
A good example for the integration of realistic inflow
conditions as well as wake interaction can be found in A standard tidal turbine requires blades, a generator and
the paper by Badoe et al. (2022), who simulated up to a support structure. One of the unique features of the
three tidal turbines under comparable complex flow con- Davidson Hill Venturi (DHV) turbine is the outer struc-
ditions representing the unique FloWave Ocean Energy ture, which contains multiple hydrofoils forming a Ven-
Research Facility (part of the University of Edinburgh). turi channel (DHV Turbines Ltd, 2022; Kirke, 2011). This
This facility provides a raisable floor for the dry installa- concept was invented by Aaron Davidson and Craig Hill
tion of not only tidal turbine models and ensures a highly of Tidal Energy Pty Ltd (Australia) and is commercialised
repeatable flow condition of up to 1.6 m/s rotatable by by DHV Turbines Ltd. The smallest cross-sectional area
360◦ owing to the circular arrangement of the flow drives. occurs at the turbine position, which augments the
W. Liu et al. (2022) validated their numerical model of incoming flow resulting in increased fluid velocities act-
a diffuser-augmented tidal turbine with a towing tank ing on the turbine. As part of the ongoing research
experiment. Feng et al. (2022) used a flume to compare work, a generalised geometry description is suggested,
the wake interaction between two ducted horizontal axis providing the current conducted parameter variation as
turbines. Obviously, the next step for the validation is the well as future additional work. The conducted numeri-
deployment of the optimised structure combined with a cal simulations are limited to a 2D-approach and focus
competitive measurement system to reduce uncertainties on the guide structure only, neglecting interaction with
in the validation. the turbine. Section 2.2 introduces the solver and basic
The presented research work uses a similar approach numerical settings. A local coordinate system is intro-
to that of Maduka and Li (2021), who investigated a duced in Section 2.3 for the geometry description, shown
ENGINEERING APPLICATIONS OF COMPUTATIONAL FLUID MECHANICS 2245

as the reference geometry in Section 2.4. A wide range attainable, as shown by Badshah et al. (2020) and Ullah
of numerical results are analysed, with primary focus et al. (2019). While a key application area of the software
on the turbine cross section (Section 2.5). Section 2.6 is the investigation and optimisation of turbines (Mulu
summarises the key aspects of the verification process, et al., 2015; Picone et al., 2021), a wide range of fur-
namely the size of the fluid domain and the mesh test. ther applications can be found (Gabl et al., 2018, 2021;
All these components are required for the variation of Lee et al., 2018; Richter et al., 2021). Recent validation
the geometry presented in Section 3, and a discussion of experiments include the optimisation of an inlet–outlet
the methodology can be found in Section 4. The opti- structure (Bermúdez et al., 2017), asymmetrical orifices
mised geometry will be investigated further numerically (Gabl et al., 2014; Gabl & Righetti, 2018) and spillway
and refined, with future experimental investigation and (Andersson et al., 2013). The high degree of validation
deployment required in order to validate the modelling within the numerical solver allows this paper to focus
data. fully on the verification as presented in Section 2.6.
The standard steady-state solver computes the Rey-
nolds-Averaged Navier–Stokes (RANS) equations, while
2.2. Numerical settings and solver
the closure problem is solved based on the Shear Stress
The geometry is generated with ANSYS SpaceClaim and Transport (SST) turbulence model (Menter, 1993, 1994;
uses the geometry parameters shown in Section 2.4. The Menter et al., 2003). This turbulence model blends a k–ω
meshing and overall handling of the files is completed in and a K– turbulence model to combine the advantages
the ANSYS-Workbench. Example files and further infor- of each specific model. The k–ω SST model is widely
mation for the geometry generation can be found in the used for a broad range of solvers (Cindori et al., 2022;
data share (Gabl et al., 2022). All simulations use the He et al., 2022; Sarkar & Savory, 2021; Yi et al., 2017).
assumption of a 2D approach, which is realised by a sin- Owing to the fact that the geometry is varied in a signif-
gle cell size in the vertical z-direction. The size of the fluid icant way and introduces different flow regimes, the auto
domain is chosen based on the verification process pre- time scale function of the solver is used with a time scale
sented in Section 2.6. A constant homogeneous velocity factor of one. For some exemplary cases, a modification
is set at the inlet face. Nominally, the inlet speed is set to 0.1 and 10 was investigated, which resulted in a slower
to 1 m/s; however, for completeness the effect of speed convergence behaviour of the numerical solution. The
variation is assessed and can be found in Section 3.9. On solver runs in all cases using double precision to reduce
the opposing side of the fluid domain, the outlet is set as the potential influence of rounding errors. Three addi-
a pressure outlet with a constant static pressure of 1 bar, tional monitor points of velocities in the cross section of
representing a typical near-shore tidal site. All other sides the turbine are used to ensure that the solution is con-
of the fluid domain box are set to symmetry with the verged and can be stopped. Further verification studies
investigated structure allocated wall (no-slip wall option can be found in Section 2.6.
with a roughness setting of ‘smooth wall’). The simulated
fluid is standard water (density = 997.0 kg·m−3 ), while
2.3. Local coordinate system
an isothermal approach is used at a constant temperature
of 10◦ C. No buoyancy is considered. A local coordinate system is defined such that the z-axis is
The presented numerical investigation uses the com- coincident with that of the vertical axis of the turbine. The
mercial Computational Fluid Dynamic (CFD) software positive x-axis is orientated along the main flow direction
ANSYS-CFX (version 2020 R2). This software was used resulting in a positive velocity. Thus the x–z-plane is the
for a wide range of numerical studies in specific aspects symmetry plane for the geometry and the y-axis aligns
of tidal turbines. Sun et al. (2022) found that a four according to the right-handed coordinate system. The
bladed vertical axis turbine brings significant advantages origin for the 2D-dimensional simulations was placed on
in the starting performance based on simulations with the bottom part of the axis.
CFX. Sun et al. (2019) focused on the fluctuation of
the angular speed and also conducted specific valida-
2.4. Reference geometry and parameters
tion experiments in the laboratory. Rehman et al. (2021)
investigated the wake behind a single horizontal axis tur- The overall geometry of the DHV turbine includes a large
bine utilising this software and Allmark et al. (2020) number of variables, with each having significant impact
integrated it into the design and blade optimisation of on the overall efficiency, production cost and mainte-
a scaled turbine. In combination with other ANSYS nance of the turbine. In order to reduce the number of
products, further capabilities to couple physics and con- parameters within the design, the first step focuses on
ducted Fluid-Structure-Interaction (FSI) simulations are the outer Venturi structure and assumes that the turbine
2246 R. GABL ET AL.

itself can be optimised in a following assessment. Simi- with


lar assumptions are made for the support structure. The
previously conducted experimental and numerical inves- − 1 · LFront ≤ xneg ≤ 0 and 0 ≤ xpos ≤ LBack
tigations ensured the principal validation of the turbine (1)
concept (Kirke, 2006, 2011) as well as a fair amount of
parameter variations. A full investigation of all parame- • The hydrofoil profiles are then located to these curved
ters and combinations would exceed the time frame of forms. For the reference geometry, a GOE 222 (MVA
any research project. Hence, a limited number of com- H.33) foil (Airfoil Tools, 2022) with a chord length
binations are investigated starting from the reference cFoil of 150 mm is chosen. The local reference point
geometry described later. Nevertheless, a key step is the for the hydrofoil is placed at the midpoint of the chord
full parametrisation of the geometry so that the results length at the maximum height (Gabl et al., 2022). This
gained later can easily be documented, reproduced and ensures that the hydrofoil does not interfere with the
expanded upon. The decision was made to provide this cross-sectional area of the turbine. Downstream of
in the form of a Python code, which is available in Gabl the central position on each side, the bell curves split
et al. (2022). This also publishes a specific IronPython into segments with a constant chord length of sTarget
version, which allows for the geometry definition in the equal to 160 mm. The implementation is shown in
3-D CAD modelling software SpaceClaim. Therewith a Gabl et al. (2022) and results in an integer number of
full integration in the ANSYS-Workbench is possible and hydrofoils for each side. As shown in Figure 1(b), the
the fluid solver ANSYS-CFX was used for this project. individual reference point for the last hydrofoil is still
For the following geometry investigation an initial ref- inside the definition point, but the hydrofoil reaches
erence geometry (Figure 1(b)) was utilised, based on a outside the curve. Limiting the design to one sin-
previous DHV demonstrator project. These values are gle size hydrofoil with constant chord spacing might
presented in combination with the specific geometry swallow smaller geometry changes but seems to be
parameters as follows. more practical than variations of the hydrofoil sizes.
Variable hydrofoil size would result in a significantly
• One of the defining parameters of the housing geom- higher cost of the structure. Furthermore, each hydro-
etry is the radius of the vertical axis turbine, RTurb . It foil is rotated based on the differential of the curve to
is chosen as 0.6 m for the reference geometry and later ensure that the tangential direction is similar to the
used to standardise all lengths of the geometry. bell shape. Additional variables that allow for a con-
• Both sides of the Venturi channel are symmetrical stant change to the angle of attack of all hydrofoils
along the x–z-plane and split at the y–z-plane (the as well as a specific modification of the first hydro-
turbine location) into two parts: the nozzle and the dif- foil (the red foil in Figure 1(b)) in the flow direction
fusor. As shown in Figure 1(b), this shape is defined are included in the Python code. The additional angle
by three points for the main curves. The center one is of the orientation of the first hydrofoil in the flow
defined by the turbine with RTurb and the coordinates direction is investigated in Figure 2.
of the outer two points are defined by RFront and LFront • In addition to the hydrofoils, a brim is included in the
as well as RBack and LBack . In the case of the reference reference geometry. The height (in the y-direction) is
geometry, shown in Figure 1(b), the combination of 100 mm and the thickness is 10 mm. These values were
[RFront ,LFront ,RBack ,LBack ] equal to [1.2, 1.1, 2.1, 1.9] · primarily chosen to be small but still clearly have an
RTurb was chosen. The radius describes the dis- influence on the mesh generation. A specific investi-
tance/length in the y-direction according to the local gation of the size of the additional brim is presented
coordinate system. in Section 3.3.
• A bell curve is defined between these points, cho-
sen based on Alkhabbaz et al. (2022), which includes The last two noteworthy aspects of the geometry
the work of Khamlaj and Rumpfkeil (2018a, 2018b). definition, namely the hydrofoils as well as the brim,
Equation (1) presents the nozzle (with a negative are clear differences from the previously tested design
x-value indicated by xneg ) and diffusor (xpos ) part of the DHV turbine. In some aspects, a smaller hydro-
aligned along the flow direction on the right side of foil is advantages compared to the larger profiles previ-
the geometry (Figure 1(b)). ously deployed. Smaller shapes can be extruded (ideal
for mass production) instead of welding single pre-cut
2
yneg = RTurb − 1 − xneg · (RFront − RTurb )/L2Front parts. In the original configuration of the DHV tur-
bine, a brim structure was absent from the design. A
2
ypos = RTurb − 1 − xpos · (RBack − RTurb )/L2Back wide range of investigations have shown the efficiency
ENGINEERING APPLICATIONS OF COMPUTATIONAL FLUID MECHANICS 2247

Figure 1. (a) Overview of the previous design provided by DHV Turbines Ltd, and (b) the half-side reference geometry and the vertical
axis turbine rotating in the origin of the coordinate system.

Figure 2. Summary of the verification process with a specific mark of the start as well as the final chosen value — (a-c) present the full
set of fluid domain changes but as a result of to the individual investigated length — (d) comparison of the different investigated meshes
αFirstFoil : (a) area-averaged velocity at the turbine cross section velc ; (b) forces in the x-direction, Fx , standardised by the reference value;
(c) kinetic energy flux coefficient, αc , for the turbine cross section; and (d) the extreme values of the velocities in the turbine section.

of such an additional blockage element for wind tur- the bell shape. Its upstream edge, closest to the symme-
bines (Alkhabbaz et al., 2022; Arumugam et al., 2021; try plane, is similarly placed to the reference points of the
Khamlaj & Rumpfkeil, 2018a, 2018b; Nunes et al., 2020). hydrofoils, but with half the chord distance sBrim equal
To test the influence in the case of a DHV turbine, a small to 80 mm. The variation of the geometry is discussed and
brim section is initially added on both sides at the end of the results are presented in Section 3.
2248 R. GABL ET AL.

2.5. Post-processing range of 1.2 (Gabl et al., 2014; Gabl & Righetti, 2018;
Ward-Smith, 1980). The α-value is calculated based on
The presented sensitivity study focuses on the Venturi
the following equation:
shaped structure and does not include the turbine, which
  
would be located in the middle of the two structures. Ekin,real 1 vel 3
This has the advantage that the computational costs can αc = = · dA (2)
Ekin,theo A A velm
be reduced, while increasing the parameter combina-
tions. But as a consequence of this limitation, a key target This equation uses an area-integral in combination with
value, namely the energy production, cannot be deliv- the local velocity in the cell vel and the area-averaged
ered directly by this numerical model. Consequently, velocity value for the investigated cross section velm . In
the post-processing of the numerical results focuses on the case of the αc -value the reference cross section is
the velocity distribution represented by the variable vel. Plane2. It can be included in ANSYS CFD-Post using the
The actual flow direction is not considered and further CFX Expression Language, or CEL, as an expression with
detailed analysis can be used for the evaluation of the the location Plane2 references the cross section at the
full geometry (Gabl et al., 2014). In the next step of turbine:
the research project, those values are correlated with the
energy production and provide a specific indication as to alphaC = areaInt(vel^3)@Plane2/
the number of runs required to find the optimum result ((areaAve(vel)@Plane2)^3*area()
for this model. @Plane2)
Three different control sections are defined. One full
cut through the fluid domain as a y–z-plane (the coor- In addition to this global evaluation value, the min-
dinate system is described in Section 2.3) with an x- imum and maximum values of the cross section are
value of −3 m for the inflow conditions and another at analysed to find the range of the flow velocities at the
x = 6 m for the outflow. The main control cross section turbine. Note that the α and extreme values are only
is set in the center of the turbine axis orientated in the shown if they are particular interesting for the analysis.
y–z-direction but limited to RTurb +10 mm on each side. Section 3 includes example analysis of velocity contour
Specifically, this allows the flow between the two struc- plots. Those plots always show the x–y-plane and the
tures at the location of the turbine to be analysed. Val- influence of the z-direction can be neglected owing to
ues referencing this cross section are indicated with the the 2D nature of the simulation. If not otherwise stated,
index ‘c ’. The other two cross sections are used for the the color bar for the velocity plots are limited to between
verification of the inflow and outflow and are not specif- zero and 3.5 m/s, which allows ease of comparison
ically reported in this paper. For the main analysis, the between the various geometry investigations provided.
following parameters are used:
2.6. Verification
• area-averaged velocity at the turbine cross section velc The verification process includes an investigation of the
using the ANSYS CFD-Post function areaAve(vel) fluid domain size based on a first pass approximation of
• total forces in the x-direction, Fx , based on the calcu- the computation mesh, which was subsequently modified
lation sum(Force X) in a second step. A compromise between the distance of
• kinetic energy flux coefficient, αc , based on the boundary conditions to the main investigated part as
Equation (2) well as the overall computational cost of the simulations is
• extreme values (minimum and maximum) of velc targeted, while ensuring that the resulting outcomes are
independent of the chosen mesh input variable. For all
The first two values, namely velc and Fx , are the main presented simulations, the reference geometry was inves-
parameters for the analysis presented in Section 3. The tigated as introduced in Section 2.4. The derived variables
average of the full cross section was deliberately used and analyses are presented in Section 2.5.
for the velocity value, while the velocity distribution An initial volume for the fluid domain was chosen
was evaluated with the help of the αc -value. This coef- based on a rectangular cuboid with three lengths (all
ficient, which is used to correct the standard Bernoulli’s referenced from the origin of the local coordinate sys-
equation for the real kinetic energy Ekin,real , is typically tem, Section 2.3) and a thickness. The nearside length
larger than the homogeneous approach for the theoreti- LFluid front defines the dimension in the negative x-
cal value Ekin,theo . Fully developed laminar flow in a pipe direction and was initially chosen to be equal 10 · RTurb ,
cross section results in a parabolic velocity distribution while both sides were extended out to an LFluid side of 10 ·
and an α-value of two. Typical turbulent flows are in the RTurb from the symmetry plane, resulting in a total width
ENGINEERING APPLICATIONS OF COMPUTATIONAL FLUID MECHANICS 2249

of 20 · RTurb . The downstream part of the fluid domain for the chosen indicator velc . The upstream section was
was expanded with a length of LFluid back = 20 · RTurb . chosen to be too short in the first assumption and was
They are highlighted as starting values in the graphs pro- hence expanded to 30 · RTurb . Similarly, the side lengths
vided in Figures 3(a)–3(c). All three graphs show the were expanded to 15 · RTurb . All dimensions were delib-
same outputs but mapped against the alternative inves- erately increased to accommodate the expansion of the
tigated lengths. To start with, the downstream section of bell shapes and any expected modifications of the geom-
the fluid domain was varied. Figure 3(a) shows that the etry. The total thickness of the fluid domain was set to the
range of 10 to 40 · RTurb results in very similar values

Figure 3. Final meshing used for the 2D-investigation using the reference geometry.

Figure 4. Overview of the main grid for the geometry variation normalised by the turbine radius RTurb with a colour map of the angle
αTip of the first (for the nozzle) or last hydrofoil using this definition point. A value of αTip = 0◦ indicates orientation along the x-axis and
90◦ alongthe y-axis. The reference value is described in Section 2.4.
2250 R. GABL ET AL.

depth of a single cell of 10 mm, thus resulting in a 2D sim- the variation of the nozzle geometry and the investi-
ulation. The thickness (fluid domain in the z-direction) gation of the diffusor is combined with the extension
was modified to 5 and 100 mm, which had no influence of the brim in Section 3.3. For all those variations, the
on the velocity distribution. radius of the vertical turbine RTurb is kept constant at
Following the variation of the fluid domain, the mesh- 600 mm. Section 3.4 provides the investigation of this
ing strategy and resolution were modified. Figure 3 parameter in two ways and complements the geome-
presents an overview of the conducted mesh test. All try variation. The narrowing of the structure’s geom-
meshes were 2D dimensional with one cell height in the etry resulted in an interesting result, which is further
z-direction. Owing to the automated geometry gener- investigated in Section 3.5 by scaling both sides of the
ation and high degree of variability, only global mesh structure with a constant factor. In addition, the angle of
parameters were set that influence the complete fluid the first hydrofoil in the flow direction is investigated in
domain. The final mesh deploys the following settings Section 3.6.
with the chosen values: element size (50 mm), growth rate The second main group of the variation is the investi-
(1.05), max size (100 mm), defeature size (0.05 mm), cur- gation of the hydrofoils themselves. Section 3.7 scales the
vature min size (0.05 mm) and curvature normal angle length of the individual hydrofoil as well as the distance
(0.5◦ ). The MultiZone method was used to fill the fluid between the standard foils. Up to this point, all simu-
domain with hexagonally shaped cells, and in total the lations are conducted with a hydrofoil profile based on
mesh included over 1.1. million nodes. Owing to the fact the GOE 222 foil (Airfoil Tools, 2022) and only within
that this meshing method is limited to a single core, mesh Section 3.8 is the hydrofoil geometry changed. Owing to
generation took a comparably long time and the deci- various factors, it was assumed that only one single type
sion was made to split the complete fluid domain into of hydrofoil was used at any time in the variation. A dis-
three sub-domains. As shown in Figure 3, each side of the cussion of the methodological approach used is provided
geometry is cut out with a total length in the x-direction in Section 4.
of 9 · RTurb (three upstream and 6 · RTurb downstream
of the turbine axis) and individual widths of 6 · RTurb .
This results in an interface boundary but allows the mesh 3.2. Variation of the nozzle
generation to be distributed to more processors. The first variation of the geometry is to alter the noz-
As a final assessment, the selected mesh was run zle of the Venturi structure. For those cases, the down-
within the fluid domain, which was expanded by the stream part is fixed with reference values, as described
factor 1.5 in the x–y-plane. For this expanded case, the in Section 2.4. The results for this variation are provided
area-averaged values of the velocities at the turbine were in Figure 5, which presents the area-averaged velocity
smaller by 0.012,77 m/s, representing 0.504% of the orig- value velc of the turbine cross section in the upper row
inal average value. and the total forces, Fx , in the x-direction. Note that the
latter is normalised by the reference value. Each graph
includes the variation of both parameters simultaneously
3. Results
and then focuses on the individual parameter. The left
3.1. Overview column presents it in relation to the standardised value
of the inlet radius RFront and the length LFront in the
The presented results can be grouped into four different
right column of Figure 5. A variation grid of 180 mm
categories: (a) the shape of the bell curve described by
(0.3 · RTurb ) was investigated for both variables defin-
the definition points; (b) the angle of the first hydrofoil;
ing the inlet bell. Furthermore, specific additional points
(c) the shape and size of the hydrofoils; and (d) the inflow
were run to investigate further specific combinations and
velocity, which is presented in Section 3.9.
enrich specific parts with a high gradient.
An overview of the major variation grid for the noz-
By varying RFront and the length LFront and comparing
zle and diffusor part is shown in Figure 4. The reference
the outcomes, the dimension in the orthogonal direc-
value, presented in Section 2.4, is marked. This figure also
tion to the main flow direction, namely RFront , is shown
provides the tangential angle of the hydrofoil to the out-
to have a significant influence. Figure 5(a) indicates that
side of the bell curve. In addition to the color bar, two
an increasing opening width RFront results in an increas-
exemplary values are marked with 10◦ and 55◦ as a ref-
ing area-averaged velocity at the turbine cross section
erence. The latter presented results also include further
velc up to a certain level, which is followed by a drop in
refinement; however, this is not included in this overview.
the efficiency of the inlet structure. This specific section
Both parts are varied separately, while the other side
is refined down to a level of 5 mm steps for RFront . The
is kept at the reference geometry. Section 3.2 presents
ENGINEERING APPLICATIONS OF COMPUTATIONAL FLUID MECHANICS 2251

Figure 5. Overview of the main grid for the geometry variation normalised by the turbine radius, RTurb , with a color map of the angle
αTip of the first (for the nozzle) or last hydrofoil using this definition point. A value of αTip = 0◦ indicates orientation along the x-axis and
90◦ along the y-axis. The reference value is described in Section 2.4.

Figure 6. Variation of the nozzle with a fixed diffusor [RBack , LBack ] = [2.1, 1.9] · RTurb : (a) area-averaged velocity at the turbine cross
section velc depending on the opening radius RFront ; and (b) sorted by the length LFront ; (c) and (d) forces in the x-direction, Fx , are
standardised by the reference value. All lengths are standardised by the turbine radius, RTurb .
2252 R. GABL ET AL.

Figure 7. Velocity plot comparing RFront = 1.5 · RTurb (left) and RFront = 1.8 · RTurb (right) with a fixed LFront = 1.1 · RTurb .

jump occurs for an RFront value in the range of 1.74583 · or even beyond the outlet RBack results in no further
RTurb (between 1045 and 1050 mm). In reality, such close improvement. Such a drop must be prevented, and the
points would not be advisable as construction tolerances final design should tolerate some uncertainty to stay in
could potentially have a significant influence of the tur- the correct flow regime.
bines operation. Therefore, a higher level can be achieved
starting with 1.5 · RTurb in order that the aforementioned
3.3. Variation of the diffusor including brim
jump can be avoided.
The total force in the x-direction, Fx , acting on the In a second step, the shape of the diffusor is varied while
structure increases for both smaller and larger values the nozzle geometry is kept constant. Smaller RBack val-
with a minimum of around RFront = 1.5 · RTurb . Conse- ues are excluded and larger steps for the variation of
quently, both values would indicate that the RFront could 360 mm (0.6 · RTurb ) are investigated. Figure 7 provides
be slightly increased. an overview similar to Figure 5, which is described in
As shown in Figure 6, a velocity drop occurs owing to Section 3.2. Similarly to the front, the parameter RBack
recirculation zone changes. A larger RFront value results has a more significant influence than the length LBack .
in the side structure acting like a blockage and reducing An increase of RBack not only directly enhances the aver-
the flow speed inside to near zero. Consequently, this is age velocity at the turbine section velc but also intensifies
not only influenced by the nozzle but also by the diffusor, the total forces acting on the structure. It has to be high-
including the brim as well as the flow speed. In the case lighted that the increase from 2.5 m/s to close to 3.25 m/s
of the larger RFront (Figure 6, right), a leakage of the flow with the larger RBack results in scaled forces, multiplied
behind the first hydrofoil is obvious, which indicates that by a factor of about eight.
this combination does not guide the flow as intended. Hence, a large RBack acts similarly to a brim; the resul-
The main conclusion, which can be drawn based on tant effects of including this additional blockage element
this variation assessment, is that the inlet dimension of are described within this section. Starting from the ref-
the reference geometry is in a satisfactory range. The erence geometry, which is presented in Section 2.4, the
parameter RFront is critical, but an expansion close to length of the additional brim is changed, the results of
ENGINEERING APPLICATIONS OF COMPUTATIONAL FLUID MECHANICS 2253

Figure 8. Variation of the diffusor with a fixed nozzle [RFront , LFront ] = [1.2, 1.1] · RTurb : (a) area-averaged velocity at the turbine cross
section velc depending on the opening radius RBack ; and (b) sorted by the length LBack ; (c) and (d) forces in the x-direction, Fx , standardised
by the reference value. All lengths are standardised by the turbine radius, RTurb .

Figure 9. Variation of the height hBrim of the additional brim: (a) averaged velocity at the turbine cross section velc ; and (b) forces in the
x-direction, Fx , standardised by the reference value. All lengths are standardised by the turbine radius, RTurb .

which are presented in Figure 8. With an initial RBack both parameters is not driven by the hydraulic charac-
of 2.1 · RTurb , the brim reaches out further to 2.4 · RTurb teristic but instead by construction costs.
and consequently covers a blockage of 4.5 · RTurb , which Figure 9 presents the extreme cases for the vari-
is similar to the maximum expansion of the geometry able RBack with the shortest length LBack . The geometry
parameter RBack of the diffusor (Figure 7). An increasing with the largest RBack introduces a massive blockage and
hBrim causes a direct increase of the indicator velc at the acts more like a brim with holes. Based on the litera-
turbine cross section but also a significant increase of the ture (Nunes et al., 2020), an inclination of the brim to the
forces Fx acting on the overall structure. The conducted front instead of to the back might be advantageous. The
variation of the hBrim indicates that the velocity converges observed convergence behaviour of the simulation with
in the range 3.2 m/s, while the forces are increased by more extreme values indicates that the flow is no longer
a factor of eight. This is very similar to the variation of steady state and a transient solver would be more appro-
the RBack and explains why both parameters are detailed priate. This is also obvious in Figure 9 (right), which
jointly within this section. Clearly, the optimisation of shows that the downstream jet is slightly orientated in the
2254 R. GABL ET AL.

Figure 10. Velocity plot comparing RBack = 1.5 · RTurb (left) and RBack = 4.5 · RTurb (right) with a fixed LBack = 1.3 · RTurb .

positive y-direction, indicating flow instability caused by The independent variation of the turbine radius, RTurb ,
the extreme values of the diffusor part of the geometry. is added more for completeness and not for their actual
An expansion of the RBack as well as the hBrim results usability within the optimisation. Since the possibility
in not only higher velocities but also increased forces. to install the turbine is not specifically checked, these
The extreme expansion of RBack can be combined or even cases are not realistic. For those simulations, the inlet
replaced with a larger brim, which might be more cost dimensions of the nozzle were fixed with RFront = 1.2 ·
effective than the hydrofoils. But for a real application, RTurb,Ref and consequently, for all RTurb values that exceed
the additional loading will be prohibitive if too large val- this length, the complete flow regime changes. Figure 10
ues are chosen. An addition of a second turbine instead shows that, for larger RTurb values, the velc is significantly
of a single large turbine should be considered in this reduced. This results in flow speeds under 1 m/s (bound-
case. ary conditions at the inlet), which would be worse than
having no supporting structure. The analysis of the αc
indicates an extreme increase of inhomogeneous flow
3.4. Variation of the turbine radius and the minimum value of the velocity falls down to
For all previously reported variations of the geometry, the 0 m/s. Such an expansion is clearly not a good choice.
turbine radius, RTurb , is set to 600 mm, which is used as Figure 11 clearly show that the recirculation zone moves
the reference value RTurb,Ref for the following variation from the outside to the inner area, which is far from a
of this parameter. Two different concepts of the variation favorable result.
are investigated: (a) the dimension RTurb is varied inde- More realistic than the previously single variation of
pendently of the other parameters (Figures 10 and 11); the RTurb is a parallel change of the structure. Where all
and (b) the change of the RTurb is similarly applied for parameters in the y-direction, namely RTurb , RFront and
the other two lengths in the y-direction of the geometry. RBack , are changed with identical values. This results in
The latter results in a parallel movement of the reference a constant offset in the y-direction keeping the geometry
geometry reducing or widening the middle cross section symmetrical to the x-axis. The results of this variation are
(Figures 12 and 13). presented in Figure 12 covering a range of a quarter of
ENGINEERING APPLICATIONS OF COMPUTATIONAL FLUID MECHANICS 2255

Figure 11. Variation of the turbine radius, RTurb , while all other geometry parameters are kept at their reference values: (a) area-averaged
velocity at the turbine cross section velc ; (b) forces in the x-direction, Fx , standardised by the reference value; (c) kinetic energy flux
coefficient, αc , for the turbine cross section; (d) and the extreme values of the velocities in the turbine cross section. All lengths are
standardised by the turbine radius of the reference geometry, RTurb,Ref = 600 mm.

Figure 12. Velocity plot comparing the reference geometry with a turbine radius, RTurb , of 400 mm, which is equal to 0.66 · RTurb,Ref (left)
in comparison with 2 · RTurb,Ref (right).
2256 R. GABL ET AL.

Figure 13. Variation of the turbine radius, RTurb , including RFront and RBack resulting in a parallel move of the sides: (a) area-averaged
velocity at the turbine cross section velc ; (b) forces in the x-direction, Fx , standardised by the reference value; (c) kinetic energy flux
coefficient, αc , for the turbine cross section; and (d) the extreme values of the velocities in the turbine cross section. All lengths are
standardised by the turbine radius of the reference geometry, RTurb,Ref = 600 mm.

the initial reference radius to the doubled value. Reducing higher velocities can generate a positive impact on energy
the space between the side Venturi structures increases production.
the average velocity at the turbine cross section velc but This section includes a theoretical investigation of a
also results in a higher total force, Fx , on the structure. single variation of the turbine radius, RTurb , that resulted
Reducing RTurb to 0.25 · RTurb,Ref increases the flow speed in some completely different flow regimes. The sec-
in the range of an additional 1 m/s while also multiplying ond variation proposes a parallel offsetting of both side
the resulting forces by a factor of 2.5 in relation to the geometries with the scaling of all three defining lengths
reference values with a RTurb,Ref = 600 mm. in the y-direction. Narrowing the channel between both
The velocity distribution becomes more homoge- sides results in increased velocities and forces. This indi-
neous with the a smaller RTurb , indicated by the αc -value cates that a combined variation of nozzle and diffusor
and a comparable small velocity difference of 0.5 m/s has the potential to improve the velocity at the turbine
(maximum subtracted by the minimum velocity). In further. How far this is beneficial for the actual energy
this case, the cross section at the turbine is reduced to production at the turbine will need to be investigated
a quarter but the discharge, which flows through the separately.
cross section, reaches approximate 34% of the discharge
through the reference geometry.
The smallest investigated distance between the two 3.5. Combination – scaling the structure
sides is compared with the largest configuration in The second part of the previous Section 3.4 shows that a
Figure 13, which shows the comparable high velocity in parallel offsetting of the full structure to the center results
the turbine cross section, resulting in a jet at the outlet. in increased speed at the center of the turbine (Figure 12).
The obvious asymmetry of the velocity distribution in the A preliminary hypothesis is the aforementioned range of
wake of the turbine indicates that an unstable flow is gen- influence that one side has over the other, which over-
erated. It seems that each side has a range of influence laps if the sides are close enough. To test this, a combined
and that, if the distance is reduced, both sides support change of the nozzle and diffusor part is investigated in
each other. To test this assumption, a specific variation this section. The factor fac is used to scale the geome-
of the full structure is presented in Section 3.5. In any try while keeping the turbine radius, RTurb , constant. This
case, this investigation has to be repeated in a full 3D factor is applied directly to the lengths LFront and LBack as
model and with a full turbine model in order to assess if well as to the value of the difference between RFront and
ENGINEERING APPLICATIONS OF COMPUTATIONAL FLUID MECHANICS 2257

Figure 14. Velocity plot comparing two cases with a parallel move of the structure and with turbine radii RTurb of 0.25·RTurb,Ref (left) and
2·RTurb,Ref (right).

RBack as presented in Equation (3): (Figure 14) the loading for the scaled version of the geom-
etry. This scaled geometry has an RBack = 3.2 · RTurb,Ref
LFront LBack RFront − RTurb and an LBack = 3.8 · RTurb,Ref . Although comparable in
fac = = =
LFront,Ref LBack,Ref RFront,Ref − RTurb geometry, with the front section fixed to the reference
RBack − RTurb geometry, the results show a velc of around 3 m/s with
= (3) higher ratios of Fx /Fx,Ref greater than four (Figure 7).
RBack,Ref − RTurb
The analysis in Figure 14 also includes a further adap-
This is obviously not perfectly scaled since the hydro- tation for the extreme values, while only the RFront and
foils remain of one type, maintaining specific lengths RBack are modified by the factor fac. For those two spe-
and distances along the channel, but it does provided cific cases, the lengths LFront and LBack are kept at their
an approximation allowing the simulation of compara- reference lengths (Section 2.4). This was completed in
ble geometry changes as conducted with the variation Sections 3.2 and 3.3, which identify the R-dimensions as
of the RTurb (Figure 12). The results of this scaling are dominant parameters. A direct comparison of the values
shown in Figure 14. A smaller factor fac down to 0.5 is for the full scaling with a factor fac of 0.5 and 2 show that
comparable to a larger RTurb (a larger distance between those specifically marked parameters result in a lower
the two sides). Both result in a reduction of the veloc- velc , a higher Fx , comparable αc -values and a wider range
ity at the turbine cross section velc and the total forces of the extreme values, which is also moved. Consequently,
Fx . A variation in the other direction shows similar- full scaling shows better results than utilising only
ity with the previously reported RTurb /RTurb,Ref = 0.5 R-scaling.
(Figure 12) a velc of approximately 3 m/s, which can also Figure 15 shows the velocity plots of the four men-
be found for a scaling factor fac = 2 in Figure 14. This tioned extreme values. The two top row pictures present
indicates that the two geometries are comparable; how- the results of the full scaling according to Equation (3).
ever, the forces show that RTurb /RTurb,Ref equates to 1.5 For the geometries in the bottom row, lengths LFront
times (Figure 12) the reference value, while it is 2.5 times and LBack are not scaled and only RFront and RBack are.
2258 R. GABL ET AL.

Figure 15. Scaling by different factors f of the structure with a fixed turbine radius, RTurb : (a) area-averaged velocity at the turbine cross
section, velc ; (b) forces in the x-direction, Fx , standardised by the reference value; (c) kinetic energy flux coefficient, αc , for the turbine
cross section; and (d) the extreme values of the velocities in the turbine cross section.

Figure 16. Velocity plot comparing four different scaled geometries: factor fac = 0.5 (left,top) and 2 (right, top) – the scaling is only
applied for the R-values with a fac = 0.5 (left, bottom) and 2 (right, bottom).
ENGINEERING APPLICATIONS OF COMPUTATIONAL FLUID MECHANICS 2259

Figure 17. Velocity plot comparing four different additional angles applied to the first hydrofoil in the flow direction: αFirstFoil = −20◦
(left,top), 20◦ (right, top), −90◦ (left, bottom) and 90◦ (right, bottom).

These results indicate that the R-lengths in the geome- The conducted variation of the parameter is delib-
try are dominating but that scaling in length along the erately expanded to investigate the range of influence
flow direction should also be considered. How far such in more detail. Extreme values of ±90◦ , for example,
an adaptation results in a real performance improvement result in a significant reduction of the velocity value velc ,
must be checked separately. while the kinetic energy flux coefficient, αc , is signifi-
cantly increased. This indicates that the first hydrofoil
introduces a disturbance that has a downstream influ-
3.6. Angle of the first hydrofoil ence. Obviously, this can also be seen in the evaluation
This specific section focuses on the orientation of the first of the extreme values in the turbine cross section shown
hydrofoil in the flow direction, describing a key function in Figure 16(d). Minimum values are reduced to 0.5 m/s,
which provides the addition of an initial direction change equal to half the inlet velocity, 1 m/s.
included to benefit the Venturi structure. The parameter Figure 16 indicates that an optimum value can be
αFirstFoil is introduced to describe this additional angle reached by including an additional angle of around −20◦ .
of this specific foil. For all other cases, the hydrofoil is The benefits are presented in the top row of Figure 17.
rotated according to the tangential direction of the bell A value of −20◦ for the parameter αFirstFoil results in
shape at the definition point of the individual hydrofoil. the trailing edge of the first hydrofoil being placed slightly
The conducted variation of this additional angle cov- closer to the symmetry plane (the x–z-plane) leading to
ers a range of ±90◦ from the original tangential orien- an overall flow velocity increase. A rotation in the oppo-
tation with αFirstFoil = 0◦ . The definition of the angle site direction generates a step and the flow detaches, the
αFirstFoil used is that a positive value results in a trail- ramification of which is an overall reduced flow speed at
ing edge rotated away from the symmetry plane. Negative the turbine cross section as well as higher forces on the
values rotate the back of the foil so that it narrows the structure. The extreme values of ±90◦ are also shown
cross section in between the two sides. Figure 16 presents in the Figure 17, representing the boundaries of these
the results, showing that incremental negative angles variations. Both cases are obviously not ideal.
improve the area-averaged velocity at the turbine cross The orientation of the first hydrofoil in the flow
section, velc occurring at velocities slightly over 2.6 m/s direction has a significant influence, while the precise
around a αFirstFoil of −20◦ . This is also connected with optimum angle is dependent on the size and type of
a reduction of the total forces in the x-direction. Con- the hydrofoil profile implemented. Consequently, this
sequently, such a modification of the first hydrofoil is parameter is one of the last criteria to optimise after key
advantageous. geometry decisions are made.
2260 R. GABL ET AL.

3.7. Hydrofoil length and distance 400 mm foil and 30 mm for the 500 mm foil. This ensures
that the individual elements are separated. Looking only
The previous variation focused on the main component
at this analysis, the average velocity at the turbine section,
of the geometry utilising a single hydrofoil GOE 222
velc , is maximised and the forces, Fx , are at a low level
(MVA H.33) (Airfoil Tools, 2022) with a chord length
(Figure 18). The four examples shown in Figure 19 indi-
cFoil of 150 mm and a linear chord distance sTarget of
cate that the choice of blade size has a significant influ-
160 mm between the definition point of each hydrofoil
ence on the structure.
on the bell curves. The type of the hydrofoil is varied in
The previous sections show that specific parameters
Section 3.8 and this section focuses on the length as well
can have a significant influence on the overall perfor-
as the distance between the foils.
mance of the turbine. It also must be noted that the
Figure 18 shows the results for different chord lengths
definition point of all cases is identical but the bell shape
of the same standard hydrofoil. A range between 50 and
can only be filled with the same hydrofoils based on a
500 mm is presented. The smaller blades use the same
constant linear chord distance along the curves. Smaller
concept for the calculation with sTarget = cFoil + 10 mm,
profiles can better reach a precise endpoint and the larger
while additional values are increased to 20 mm for the
profiles would have to be spaced perfectly to reach a

Figure 18. Variation of the chord length, cFoil , of the hydrofoil: (a) area-averaged velocity at the turbine cross section, velc ; and (b) forces
in the x-direction, Fx , standardised by the reference value.

Figure 19. Velocity plot comparing four different lengths of the hydrofoil: 50 mm(left,top), 200 mm (left, bottom), 300 mm (right,
bottom) and 500 mm (right, top).
ENGINEERING APPLICATIONS OF COMPUTATIONAL FLUID MECHANICS 2261

Figure 20. Variation of the length sTarget , which defines the distance between the reference points on the bell shape: (a) area-averaged
velocity at the turbine cross section, velc ; and (b) forces in the x-direction, Fx , standardised by the reference value.

Figure 21. Velocity plot comparing two different chord lengths, sTarget , between the definition points of the hydrofoils: 166 mm (left)
and 300 mm (right).

better comparability. Keeping the limitation of this com- of velc for the reference geometry. Consequently, for this
parison in mind, all investigated lengths performed in case, two variation points are reported in Figure 20 for
a very comparable way and might need more detailed the sTarget length of 160 mm. A slight expansion results in
optimisation to reach an optimum. an overall improvement of the velc , and a reduction of the
In a second step, the linear chord distance sTarget is total forces, Fx , can be observed, with an optimum point
varied to investigate the initial choice of 160 mm. This is at 166 mm. It is not surprising that larger values intro-
conducted for the hydrofoil GOE 222 (MVA H.33) (Air- duce large gaps in the structure and hence the efficiency
foil Tools, 2022) only, with a chord length, cFoil , set to is reduced. Figure 21 shows a comparison of a smallest
150 mm. sTarget is varied between 155 and 300 mm, which value of sTarget of 166 mm and a largest value of 300 mm.
is double the chord length of the hydrofoil. For this vari- With a larger gap between the individual hydrofoils, the
ation, the step size s is changed from 10 to 1 mm in behaviour of the structure changes from joint guidance
order to represent the small alterations of the parame- of the flow into separate obstacles in the flow.
ters within each simulation better. This value is used to This variation of hydrofoil chord length and the
find the required x-distances for the following definition definition point indicates that the initially chosen values
point, indicating that these changes result in a reduction are in a good range. Nevertheless, the distance between
2262 R. GABL ET AL.

Figure 22. Overview of the investigated hydrofoils.

the definition points of the hydrofoils indicates some The results of the variation of the hydrofoil types is
optimisation potential. However, extremes in hydrofoil presented in Figure 23, sorted depending on the area-
sizing do not provide better results. The crucial point for averaged velocity at the turbine cross section, velc . A
this decision is more likely the availability and costs of the slight improvement in velocity as well as the total forces
production of the hydrofoils. in the x-direction is noted with the NACA 2418 hydro-
foil. All other foils resulted in worse conditions with a
massive reduction for the circular cross section. The lat-
ter requires a much more massive structure while only
3.8. Hydrofoil type
producing a slightly higher total force. The kinetic energy
In this section, the type of hydrofoil is altered, while keep- flux coefficient, αc , is very similar for all investigated
ing a constant chord length, cFoil = 150 mm, and distance hydrofoils with the largest inhomogeneity shown for the
between the definition points for the individual foils, circular profile, which provides the most extreme values
sTarget = 160 mm. Similar to GOE 222, the geometry data for the velocities in the cross section.
is sourced from the online database Airfoil Tools (2022), Figure 24 shows a comparison of four of the eight
which provides over 1600 different foil types. The selec- investigated options. The top row shows the best and
tion of additional foils is conducted based on random worst options side by side and it is obvious that the circu-
choice, while trying to cover a wider range of different lar cross section is not a good option for this approach.
shapes. In addition, a circular cross section is also tested, The comparable large radius of the circle in the lateral
for which the sTarget has to be increased to 165 mm. (Note direction results in a comparable massive structure and
that, for this simulation, the definition points are located it works more like a solid Venturi channel with addi-
on the inside of the turbine and not on the centerline tional wall roughness. The tip of the foil of the very thin
of the hydrofoils.) Figure 22 provides an overview of the M23 hydrofoil is remarkable in comparison to the other
investigated types. hydrofoils. In this case, the flow detaches very early and
ENGINEERING APPLICATIONS OF COMPUTATIONAL FLUID MECHANICS 2263

Figure 23. Variation of the hydrofoil types: (a) area-averaged velocity at the turbine cross section, velc ; (b) forces in the x-direction, Fx ,
standardised by the reference value; (c) kinetic energy flux coefficient, αc , for the turbine cross section; and (d) extreme values of the
velocities in the turbine cross section.

results in a small recirculation zone behind the first foil. for the generation of the geometry in Gabl et al. (2022)
On the other hand, the two foils on the left cause a veloc- and a variation is presented in Section 3.6.
ity peak at this location. This indicates that the fixed angle
along the bell curve with an orientation along the tangen-
tial direction is not ideal for all foils. A correction of the 3.9. Inlet velocity
first foil in the incoming flow direction could result in an All previous simulations assumed that the inlet velocity is
improvement of the results for this specific foil. constant at 1 m/s, which is introduced homogeneously at
This section analyses seven additional hydrofoils for the inlet boundary conditions. In this section, the value
comparison with the originally used GOE 222 foil. A is varied over a range of 0.5 to 2 m/s, while keeping all
slight improvement can be found by changing to NACA the other assumptions the same, including the constant
2418, which is more symmetric and has the potential to reference geometry (Section 2.4). Figure 25 presents the
be less expensive to manufacture. The distance between results of this variation. Contrary to the previous anal-
the definition points of the hydrofoils may allow for fur- ysis, the velocity at the turbine cross section, velc , is
ther improvements, such as that shown in Section 3.7. normalised by the inlet velocity vInlet . Hence the stan-
Importantly, it is found that the orientation of the first dard vInlet is 1 m/s, the values can be compared directly
foil in the flow direction is critical, and hence a correction with the previous results, being non-dimensionalised.
angle for this specific profile is added in the Python code This value changes in relation to vInlet in a range of
2264 R. GABL ET AL.

Figure 24. Velocity plot comparing the reference geometry with three other types of hydrofoil: NACA 2418 (left,top), GOE 222 (left,
bottom), M23 (right, bottom) and circular (right, top).

Figure 25. (a) Variation of the inlet velocity, vInlet , to investigate the standardised velocity at the turbine cross section, velc ; and (b) the
total forces, Fx , on the structure in the x-direction.

approximately ±0.2 from the reference value. A doubling making it necessary to optimise the specific design for
of the vInlet causes an increase of the forces by a factor each deployment site. Extreme flow speeds also have to
of approximately 3.5. Tidal flows are generally very pre- be considered in the design of the support structure to
dictable but nevertheless extreme flow speeds can occur ensure the structural integrity of the device.
resulting in extreme loads on the structure.
Figure 26 provides the slowest and fastest investigated
flow speeds of 0.5 and 2 m/s, respectively. A separate color 4. Discussion
bar is used for each result, with maximum values scaled
The obvious limitation of the presented work is that the
provided according to the inlet velocity. The downstream
turbine itself is not simulated, which allows the param-
section with close to 0 m/s is smaller for the increased
eters as well as the overall needed computational cost
flow speed. A higher gradient in the cross section can be
to be reduced. In the next steps of the research project,
observed for the higher speed, which can also be seen in
the complexity of the numerical model will be expanded,
an increased normalised velc /vInlet value.
while the needed individual runs can be reduced based
In summary, the change of inlet velocity has an influ-
on an improved understanding of the outer Venturi
ence on the velocity at the turbine cross section, velc ,
structure. Another simplification is that the numerical
ENGINEERING APPLICATIONS OF COMPUTATIONAL FLUID MECHANICS 2265

Figure 26. Variation of the inlet velocity, vInlet , with 0.5 m/s (a) and 2 m/s (b) – color bar scaled with the same factor.

simulations are limited to a 2D approach. All the com- designated refinement close to the wall should be consid-
putation meshes used have only one cell in the vertical ered. A verification process is conducted and presented
direction, hence assuming that there is no change in in Section 2.6, but it is limited to the reference geometry.
the z-direction. This neglects potential velocity differ- It is assumed that the mesh independence of the results is
ences, which are significant closer to the ground (Ahmed also a given for changes in the geometry. Section 3 shows
et al., 2017; Badshah et al., 2018; Ke et al., 2020). Further- that some variations of the geometry cause changes in
more, the structure has clear top and bottom parts, which the location of the recirculating zones and the concen-
is a plane cross section in the previous tests (Figure 1(a)). tration of the wake in a jet. For those cases, an additional
This causes additional interactions and has to be included check of the fluid domain and the mesh resolution would
in the overall optimisation. be advisable. Nevertheless, those geometries are not likely
It is assumed that the current direction is perfectly to be chosen in the future as their performance is not an
aligned, homogeneously distributed and constant. Obvi- improvement.
ously, this is an idealisation and, even in deep current The individual parameters were varied based on an
streams, the interaction of waves as well as large turbu- initial grid and in some cases further combinations were
lence in the flow cause variations in the velocity distribu- added to investigate interesting aspects. Those refine-
tion or changes in the flow directions. In all cases, only ments were triggered manually, based on the analysis,
the steady-state solver is used and some results indicate and allowed the sensitivities of the individual parameters
that, even with a fully perfect inflow condition, the result- to be identified. The investigation of this many parame-
ing wake can be an unsteady flow. The following detailed ters and also their possible combinations would be more
optimisation has to consider the usage of a fully transient efficiently handled with a specific software tool such as
solver. The k–ω SST model chosen is a standard turbu- ANSYS optiSLang. This software allows a very robust
lence model for comparable applications, but under spe- and cost-effective parameter combination to be gener-
cific conditions other approaches might be more accurate ated with (optimised) Latin Hypercube Sampling and
for hydrofoils. This should be investigated as part of the calculates a metamodel of optimal prognosis to reduce
final validation. the needed iterations to reach an optimum solution. The
The computational grid is a standard point, which usage of such an additional tool will allow the effort for
can be always improved. All current simulations use a the next optimisation steps to be reduced.
comparable fine mesh around the profiles but not a des- Further limitations can be found in the geometry
ignated inflation layer. Consequently, the y+ -values are definition. All cases are built using a single hydrofoil and
comparable high, but similar to comparable investiga- with an identical chord distance set on the bell curves.
tions (Maduka & Li, 2021) for a further optimisation a Hence, only full hydrofoils are added, and not all changes
2266 R. GABL ET AL.

in the geometry definition have an immediate impact resulting in extreme operational forces. It can be con-
in the actual geometry. A certain threshold has to be cluded that the hydrofoil chord length does not signifi-
exceeded to trigger the addition or removal of a pro- cantly influence operational criteria, but that fabrication
file. Nevertheless, the conducted geometry generation is costs would need to be considered as part of an over-
ideal for such an exploitative approach but can be refined all device optimisation. By increasing the gap between
in an additional detailed optimisation. Furthermore, the the hydrofoils as well as exchanging the GOE 222 with
hydrofoils are always orientated according to the tan- an NACA 2418 hydrofoil, some improvements in output
gential direction. All of these aspects can be individu- were shown. In addition to the variations of the geometry,
ally modified and varied for each hydrofoil. Obviously, the inlet velocity was varied showing that better results
the number of potential options for the investigations can be achieved with higher flow speeds, but no signifi-
increases dramatically. cant decrease for lower speeds. Overall, these variations
The research presented focuses on the Venturi struc- show that the chosen reference geometry results provide
ture and neglects the influence of the vertical axis tur- good system performance, with only small improvements
bine. In subsequent work, the improvements found will being achieved by tweaking design constraints. The next
be checked to see if the changes are really beneficial step for the research is to expand the results from the 2D
for overall energy production. Currently, only fully sym- approach and integrate a full supporting structure as well
metrical geometries have been investigated, but there as turbine optimisation.
may be benefits in modifying one side, allowing for
optimisation of the geometry where the rotating tur-
bine moves against the flow direction. Further detailed Disclosure statement
investigations are needed to improve the overall effi- Mark Hill is the owner of DHV Turbines Ltd, which com-
ciency of the Venturi shaped structure. The experimen- mercialises the tidal turbine concept presented in this
tal investigation of the finally optimised geometry is work. The authors declare no further conflict of interest.
desirable, which would validate the numerical simula-
tion and provide additional trust in the numerical sim-
Funding
ulation to cover site-specific adaptations of the turbine
geometry. This work was supported by Interreg North-West Europe
(NWE) program and the work was carried out as part of the
Marine Energy Alliance (MEA) project.

5. Conclusions
Data availability statement
The research presented focuses on the hydraulic perfor- If there is a data set associated with the paper, please provide
mance of the surrounding structure of the Davidson Hill information about where the data supporting the results or
Venturi (DHV) tidal turbine. Multiple hydrofoils were analyses presented in the paper can be found. Where applicable,
placed in a Venturi shape to increase the flow speed at the this should include a hyperlink, DOI or other persistent identi-
fier associated with the data set(s). Templates are also available
vertical axis turbine. The first step was to provide a gen-
to support authors.
eralised description of the structure (Gabl et al., 2022),
which was done in Python allowing future expansions,
including using a different numerical solver. In the sec- ORCID
ond step, this geometry description was adapted for the Roman Gabl http://orcid.org/0000-0001-9701-879X
ANSYS-Workbench with the commercial solver ANSYS- David M. Ingram http://orcid.org/0000-0002-8669-8942
CFX, the files for which are available in the connected
datashare (Gabl et al., 2022). A wide range of geometry References
parameters were investigated in a 2D approach without
Ahmed, U., Apsley, D., Afgan, I., Stallard, T., & Stansby, P.
the turbine, based on a reference geometry (Section 2.4). (2017). Fluctuating loads on a tidal turbine due to velocity
This variation included the nozzle part as well as the dif- shear and turbulence: Comparison of CFD with field data.
fusor, including the brim and an additional angle of the Renewable Energy, 112, 235–246. https://doi.org/10.1016/j.
first hydrofoil in the flow direction. In general, the expan- renene.2017.05.048
sion of the investigated structures resulted in an increased Airfoil Tools (2022). Online database for the airfoil GOE 222.
[2022-05-18] http://airfoiltools.com/airfoil/details?airfoil =
area-averaged velocity at the turbine cross section; how-
goe222-il
ever, the forces on the structure were also increased. The Alidadi, M., Klaptocz, V., Rawlings, G. W., Nabavi, Y., &
variation of the turbine radius on both sides simultane- Calisal, S. (2011). A duct optimization study for a verti-
ously showed potential for further improvements without cal axis hydro-current turbine model. Journal of Offshore
ENGINEERING APPLICATIONS OF COMPUTATIONAL FLUID MECHANICS 2267

Mechanics and Arctic Engineering, 134(2), Article 021201. CFD-embedded BEM. Renewable Energy, 108, 622–634.
https://doi.org/10.1115/1.4004955 https://doi.org/10.1016/j.renene.2016.10.048
Alidadi, M., Nabavi, Y., & Çalışal, S. (2008). The effect of Bermúdez, M., Cea, L., Puertas, J., Conde, A., Martín, A., &
ducting on the performance of a vertical axis tidal turbine. Baztán, J. (2017). Hydraulic model study of the intake–outlet
In Proceedings of the ASME 2008 27th international confer- of a pumped-storage hydropower plant. Engineering Appli-
ence on offshore mechanics and arctic engineering (Vol. 6, cations of Computational Fluid Mechanics, 11(1), 483–495.
pp. 825–830). American Society of Mechanical Engineers. https://doi.org/10.1080/19942060.2017.1314869
https://doi.org/10.1115/OMAE2008-57902 Borg, M. G., Xiao, Q., Allsop, S., Incecik, A., & Peyrard,
Alkhabbaz, A., Yang, H. S.., Tongphong, W., & Lee, Y. H.. C. (2020). A numerical performance analysis of a ducted,
(2022). Impact of compact diffuser shroud on wind turbine high-solidity tidal turbine. Renewable Energy, 159, 663–682.
aerodynamic performance: CFD and experimental investi- https://doi.org/10.1016/j.renene.2020.04.005
gations. International Journal of Mechanical Sciences, 216, Borg, M. G., Xiao, Q., Allsop, S., Incecik, A., & Peyrard, C.
Article 106978. https://doi.org/10.1016/j.ijmecsci.2021.10 (2021). A numerical structural analysis of ducted, high-
6978 solidity, fibre-composite tidal turbine rotor configurations in
Allmark, M., Ellis, R., Lloyd, C., Ordonez-Sanchez, S., Johan- real flow conditions. Ocean Engineering, 233, Article 109087.
nesen, K., Byrne, C., Johnstone, C., O’Doherty, T., & Mason- https://doi.org/10.1016/j.oceaneng.2021.109087
Jones, A. (2020). The development, design and character- Cindori, M., Čajić, P., Džijan, I., Juretić, F., & Kozmar, H.
isation of a scale model Horizontal Axis Tidal Turbine (2022). A comparison of major steady RANS approaches
for dynamic load quantification. Renewable Energy, 156, to engineering ABL simulations. Journal of Wind Engi-
913–930. https://doi.org/10.1016/j.renene.2020.04.060 neering and Industrial Aerodynamics, 221, Article 104867.
Andersson, A. G., Andreasson, P., & Lundström, T. S. (2013). https://doi.org/10.1016/j.jweia.2021.104867
CFD-modelling and validation of free surface flow during DHV Turbines Ltd (2022). Company homepage. [2022-05-18]
spilling of reservoir in down-scale model. Engineering Appli- http://www.dhvturbines.com
cations of Computational Fluid Mechanics, 7(1), 159–167. El-Sawy, M., Shehata, A. S., Elbatran, A. A., & Tawfiq, A. (2022).
https://doi.org/10.1080/19942060.2013.11015461 Numerical simulation of flow in hydrokinetic turbine chan-
Arumugam, P., Ramalingam, V., & Bhaganagar, K. (2021). A nel to improve its efficiency by using first- and second-law
pathway towards sustainable development of small capacity efficiency analysis. Ocean Engineering, 244, Article 110400.
horizontal axis wind turbines – Identification of influencing https://doi.org/10.1016/j.oceaneng.2021.110400
design parameters and their role on performance analysis. Feng, B., Liu, X., Ying, Y., Si, Y., Zhang, D., & Qian, P.
Sustainable Energy Technologies and Assessments, 44, Article (2022). Research on the tandem arrangement of the ducted
101019. https://doi.org/10.1016/j.seta.2021.101019 horizontal-axis tidal turbine. Energy Conversion and Man-
Badoe, C. E., Edmunds, M., Williams, A. J., Nambiar, A., Sellar, agement, 258, Article 115546. https://doi.org/10.1016/j.enco
B., Kiprakis, A., & Masters, I. (2022). Robust validation of a nman.2022.115546
generalised actuator disk CFD model for tidal turbine analy- Fleming, C. F., & Willden, R. H. (2016). Analysis of bi-
sis using the FloWave ocean energy research facility. Renew- directional ducted tidal turbine performance. International
able Energy, 190, 232–250. https://doi.org/10.1016/j.renene. Journal of Marine Energy, 16, 162–173. https://doi.org/10.10
2022.03.109 16/j.ijome.2016.07.003
Badshah, M., Badshah, S., & Jan, S. (2020). Comparison of Gabl, R., Achleitner, S., Neuner, J., & Aufleger, M. (2014). Accu-
computational fluid dynamics and fluid structure interac- racy analysis of a physical scale model using the example
tion models for the performance prediction of tidal current of an asymmetric orifice. Flow Measurement and Instrumen-
turbines. Journal of Ocean Engineering and Science, 5(2), tation, 36, 36–46. https://doi.org/10.1016/j.flowmeasinst.20
164–172. https://doi.org/10.1016/j.renene.2022.03.109 14.02.001
Badshah, M., Badshah, S., & Kadir, K. (2018). Fluid struc- Gabl, R., Burchell, J., Hill, M., & Ingram, D. (2022). Geome-
ture interaction modelling of tidal turbine performance and try generation for DHV tidal turbine – Basic Python code
structural loads in a velocity shear environment. Energies, and implementation in ANSYS-CFX and Rhinoceros. Edin-
11(7). https://doi.org/10.3390/en11071837 burgh DataShare. [2022-05-18]. https://doi.org/10.7488/ds/
Bahaj, A. S. (2013). Marine current energy conversion: The 3455
dawn of a new era in electricity production. Philosophical Gabl, R., Gems, B., Birkner, F., Hofer, B., & Aufleger,
Transactions of the Royal Society A: Mathematical, Physi- M. (2018). Adaptation of an existing intake structure
cal and Engineering Sciences, 371(1985), Article 20120500. caused by increased sediment level. Water, 10(8), 1066.
https://doi.org/10.1098/rsta.2012.0500 https://doi.org/10.3390/w10081066
Baratchi, F., Jeans, T., & Gerber, A. (2020). Assessment of Gabl, R., & Righetti, M. (2018). Design criteria for a type of
blade element actuator disk method for simulations of asymmetric orifice in a surge tank using CFD. Engineer-
ducted tidal turbines. Renewable Energy, 154, 290–304. ing Applications of Computational Fluid Mechanics, 12(1),
https://doi.org/10.1016/j.renene.2020.02.098 397–410. https://doi.org/10.1080/19942060.2018.1443837
Behrouzi, F., Nakisa, M., Maimun, A., & Ahmed, Y. M. Gabl, R., Wippersberger, M., Seibl, J., Kröner, C., & Gems, B.
(2016). Global renewable energy and its potential in (2021). Submerged wall instead of a penstock shutoff valve
Malaysia: A review of Hydrokinetic turbine technology. – alternative protection as part of a refurbishment. Water,
Renewable and Sustainable Energy Reviews, 62, 1270–1281. 13(16), 2247. https://doi.org/10.3390/w13162247
https://doi.org/10.1016/j.rser.2016.05.020 Gu, Y., Lin, Y., Xu, Q., Liu, H., & Li, W. (2018). Blade-pitch
Belloni, C., Willden, R., & Houlsby, G. (2017). An investiga- system for tidal current turbines with reduced variation
tion of ducted and open-centre tidal turbines employing pitch control strategy based on tidal current velocity preview.
2268 R. GABL ET AL.

Renewable Energy, 115, 149–158. https://doi.org/10.1016/j. Li, Y., & Çalışal, S. M. (2010b). Three-dimensional effects and
renene.2017.07.034 arm effects on modeling a vertical axis tidal current turbine.
Hardisty, J. (2008). Modelling and testing the vertical axis, Renewable Energy, 35(10), 2325–2334. https://doi.org/10.10
impulse rotor tidal power pontoon. Journal of Marine Engi- 16/j.renene.2010.03.002
neering & Technology, 7(1), 1–9. https://doi.org/10.1080/204 Liu, H., Li, Y., Lin, Y., Li, W., & Gu, Y. (2020). Load reduc-
64177.2008.11020207 tion for two-bladed horizontal-axis tidal current turbines
He, J w., Zhao, W w., Wan, D c., & Wang, Y q. (2022). Numeri- based on individual pitch control. Ocean Engineering, 207,
cal study of free end effect of cylinder with low aspect ratios Article 107183. https://doi.org/10.1016/j.oceaneng.2020.
on vortex induced motion. Journal of Hydrodynamics, 34(1), 107183
106–115. https://doi.org/10.1007/s42241-022-0011-x Liu, W., Liu, L., Wu, H., Chen, Y., Zheng, X., Li, N., &
Huang, B., Gong, Y., Wu, R., Wang, P., Chen, J., & Wu, P. (2022). Zhang, Z. (2022). Performance analysis and offshore appli-
Study on hydrodynamic performance of a horizontal axis cations of the diffuser augmented tidal turbines. Ships
tidal turbine with a lobed ejector. Ocean Engineering, 248, and Offshore Structures, 10pp. Advance online publication.
Article 110769. https://doi.org/10.1016/j.oceaneng.2022. https://doi.org/10.1080/17445302.2022.2027691
110769 Liu, Z., Wang, Z m., Shi, H d., & Qu, H l. (2019). Numerical
Ke, S., Wen-Quan, W., & Yan, Y. (2020). The hydrodynamic per- study of a guide-vane-augmented vertical darrieus tidal-
formance of a tidal-stream turbine in shear flow. Ocean Engi- current-turbine. Journal of Hydrodynamics, 31(3), 522–530.
neering, 199, Article 107035. https://doi.org/10.1016/j.ocean https://doi.org/10.1007/s42241-018-0117-3
eng.2020.107035 Maduka, M., & Li, C. (2021). Numerical study of ducted
Khamlaj, T. A., & Rumpfkeil, M. (2018b). Optimization study turbines in bi-directional tidal flows. Engineering Applica-
of shrouded horizontal axis wind turbine. In Proceedings of tions of Computational Fluid Mechanics, 15(1), 194–209.
the 2018 wind energy symposium. Paper No. AIAA 2018- https://doi.org/10.1080/19942060.2021.1872706
0996. American Institute of Aeronautics and Astronautics. Menter, F. R. (1993). Zonal two equation k–w turbulence
https://doi.org/10.2514/6.2018-0996 models for aerodynamic flows. In Proceedings of the 23rd
Khamlaj, T. A., & Rumpfkeil, M. P. (2018a). Analysis and opti- fluid dynamics, plasmadynamics, and lasers conference. Paper
mization of ducted wind turbines. Energy, 162, 1234–1252. No. AIAA-93-2906. American Institute of Aeronautics and
https://doi.org/10.1016/j.energy.2018.08.106 Astronautics. https://doi.org/10.2514/6.1993-2906
Khan, M., Bhuyan, G., Iqbal, M., & Quaicoe, J. (2009). Hydroki- Menter, F. R. (1994). Two-equation eddy-viscosity turbulence
netic energy conversion systems and assessment of hori- models for engineering applications. AIAA Journal, 32(8),
zontal and vertical axis turbines for river and tidal appli- 1598–1605. https://doi.org/10.2514/3.12149
cations: A technology status review. Applied Energy, 86(10), Menter, F. R., Kuntz, M., & Langtry, R. (2003). Ten years of
1823–1835. https://doi.org/10.1016/j.apenergy.2009.02.017 industrial experience with the SST turbulence model. Tur-
Khojasteh, D., Lewis, M., Tavakoli, S., Farzadkhoo, M., bulence, Heat and Mass Transfer, 4(1), 625–632.
Felder, S., Iglesias, G., & Glamore, W. (2022). Sea level Modali, P. K., Vinod, A., & Banerjee, A. (2021). Towards a bet-
rise will change estuarine tidal energy: A review. Renew- ter understanding of yawed turbine wake for efficient wake
able and Sustainable Energy Reviews, 156, Article 111855. steering in tidal arrays. Renewable Energy, 177, 482–494.
https://doi.org/10.1016/j.rser.2021.111855 https://doi.org/10.1016/j.renene.2021.05.152
Kirke, B. (2006). Developments in ducted water current tur- Mosbahi, M., Elgasri, S., Lajnef, M., Mosbahi, B., & Driss,
bines. Tidal paper 25-04-06. Tethys Engineering. https://teth Z. (2021). Performance enhancement of a twisted Savonius
ys-engineering.pnnl.gov/sites/default/files/publications/Kir hydrokinetic turbine with an upstream deflector. Interna-
ke-2006.pdf tional Journal of Green Energy, 18(1), 51–65. https://doi.org/
Kirke, B. (2011). Tests on ducted and bare helical and straight 10.1080/15435075.2020.1825444
blade Darrieus hydrokinetic turbines. Renewable Energy, Mulu, B. G., Cervantes, M. J., Devals, C., Vu, T. C.,
36(11), 3013–3022. https://doi.org/10.1016/j.renene.2011. & Guibault, F. (2015). Simulation-based investigation of
03.036 unsteady flow in near-hub region of a Kaplan Tur-
Klaptocz, V., Rawlings, G., Nabavi, Y., Alidadi, M., Li, Y., & bine with experimental comparison. Engineering Applica-
Çalışal, S. (2007). Numerical and experimental investigation tions of Computational Fluid Mechanics, 9(1), 139–156.
of a ducted vertical axis tidal current turbine. In Proceed- https://doi.org/10.1080/19942060.2015.1004816
ings of the 7th European wave and tidal energy conference Nachtane, M., Tarfaoui, M., Goda, I., & Rouway, M. (2020).
(EWTEC 2007). European Tidal and Wave Energy Confer- A review on the technologies, design considerations and
ence/Tethys Engineering. https://tethys-engineering.pnnl.g numerical models of tidal current turbines. Renewable
ov/publications/numerical-experimental-investigation-duct Energy, 157, 1274–1288. https://doi.org/10.1016/j.renene.20
ed-vertical-axis-tidal-current-turbine 20.04.155
Lee, S. O., Seong, H., & Kang, J. W. (2018). Flow-induced vibra- Noorollahi, Y., Ghanbari, S., & Tahani, M. (2020). Numeri-
tion of a radial gate at various opening heights. Engineer- cal analysis of a small ducted wind turbine for performance
ing Applications of Computational Fluid Mechanics, 12(1), improvement. International Journal of Sustainable Energy,
567–583. https://doi.org/10.1080/19942060.2018.1479662 39(3), 290–307. https://doi.org/10.1080/14786451.2019.168
Li, Y., & Çalışal, S. M. (2010a). A discrete vortex method 5520
for simulating a stand-alone tidal-current turbine: Model- Nunes, M. M., Brasil Junior, A. C., & Oliveira, T. F. (2020). Sys-
ing and validation. Journal of Offshore Mechanics and Arctic tematic review of diffuser-augmented horizontal-axis tur-
Engineering, 132(3), Article 031102. https://doi.org/10.1115/ bines. Renewable and Sustainable Energy Reviews, 133, Arti-
1.4000499 cle 110075. https://doi.org/10.1016/j.rser.2020.110075
ENGINEERING APPLICATIONS OF COMPUTATIONAL FLUID MECHANICS 2269

Patel, V. K., & Patel, R. S. (2022). Optimization of an angle speed fluctuation. Ships and Offshore Structures, 14(sup1),
between the deflector plates and its orientation to enhance 311–319. https://doi.org/10.1080/17445302.2019.1589975
the energy efficiency of Savonius hydrokinetic turbine for Sun, K., Yi, Y., Zhang, J., Zhang, J., Haider Zaidi, S. S., & Sun, S.
dual rotor configuration. International Journal of Green (2022). Influence of blade numbers on start-up performance
Energy, 19(5), 476–489. https://doi.org/10.1080/15435075.20 of vertical axis tidal current turbines. Ocean Engineering,
21.1947821 243, Article 110314. https://doi.org/10.1016/j.oceaneng.2021
Picone, C., Sinagra, M., Aricó, C., & Tucciarelli, T. (2021). .110314
Numerical analysis of a new cross-flow type hydraulic tur- Ullah, H., Hussain, M., Abbas, N., Ahmad, H., Amer,
bine for high head and low flow rate. Engineering Applica- M., & Noman, M. (2019). Numerical investigation of
tions of Computational Fluid Mechanics, 15(1), 1491–1507. modal and fatigue performance of a horizontal axis tidal
https://doi.org/10.1080/19942060.2021.1974559 current turbine using fluid–structure interaction. Jour-
Pisetta, G., Le Mestre, R., & Viola, I. M. (2022). Morphing nal of Ocean Engineering and Science, 4(4), 328–337.
blades for tidal turbines: A theoretical study. Renewable https://doi.org/10.1016/j.joes.2019.05.008
Energy, 183, 802–819. https://doi.org/10.1016/j.renene.2021 Walker, S., & Thies, P. (2021). A review of component and
.10.085 system reliability in tidal turbine deployments. Renew-
Rehman, Z. U., Badshah, S., Rafique, A. F., Badshah, M., Jan, S., able and Sustainable Energy Reviews, 151, Article 111495.
& Amjad, M. (2021). Effect of a support tower on the perfor- https://doi.org/10.1016/j.rser.2021.111495
mance and wake of a tidal current turbine. Energies, 14(4). Wang, H.-M., Qu, X.-K., Chen, L., Tu, L.-Q., & Wu, Q.-
https://doi.org/10.3390/en14041059 R. (2020). Numerical study on energy-converging effi-
Richter, W., Vereide, K., Mauko, G., Havrevoll, O. H., Schnei- ciency of the ducts of vertical axis tidal current turbine
der, J., & Zenz, G. (2021). Retrofitting of pressurized in restricted water. Ocean Engineering, 210, Article 107320.
sand traps in hydropower plants. Water, 13(18), 2515. https://doi.org/10.1016/j.oceaneng.2020.107320
https://doi.org/10.3390/w13182515 Wang, S.-q., Xu, G., Zhu, R.-q., & Wang, K. (2018). Hydro-
Roshanmanesh, S., Hayati, F., & Papaelias, M. (2020). Chap- dynamic analysis of vertical-axis tidal current turbine with
ter 10 – Tidal turbines. In M. Papaelias, F.P.G. Márquez, & surging and yawing coupled motions. Ocean Engineer-
A. Karyotakis (Eds.), Non-destructive testing and condition ing, 155, 42–54. https://doi.org/10.1016/j.oceaneng.2018.02.
monitoring techniques for renewable energy industrial assets 044
(pp. 143–158). Butterworth–Heinemann. https://doi.org/10. Wang, Z., Qu, F., Wang, Y., Luan, Y., & Wang, M. (2021).
1016/B978-0-08-101094-5.00010-1 Research on the lean and swept optimization of a single stage
Samadi, M., Ghodsi Hassanabad, M., & Mozafari, S. B. (2022). axial compressor. Engineering Applications of Computational
Performance enhancement of low speed current Savonius Fluid Mechanics, 15(1), 142–163. https://doi.org/10.1080/199
tidal turbines through adding semi-cylindrical deflectors. 42060.2020.1862708
Ocean Engineering, 259, Article 111873. https://doi.org/10.10 Ward-Smith, A. J. (1980). Internal fluid flow – the fluid dynamics
16/j.oceaneng.2022.111873 of flow in pipes and ducts (Vol. 81). Clarendon Press.
Sarkar, D., & Savory, E. (2021). Numerical modeling of Yi, P., Wang, Y., Sun, X., Huang, D., & Zheng, Z. (2017).
freestream turbulence decay using different commercial The effect of variations in first- and second-order deriva-
computational fluid dynamics codes. Journal of Fluids Engi- tives on airfoil aerodynamic performance. Engineering Appli-
neering, 143(4), Article 041503. https://doi.org/10.1115/1.40 cations of Computational Fluid Mechanics, 11(1), 54–68.
49679 https://doi.org/10.1080/19942060.2016.1246264
Sun, K., Ji, R., Zhang, J., Li, Y., & Wang, B. (2021). Investigations Zhu, D., Xiao, R., Yao, Z., Yang, W., & Liu, W. (2020). Optimiza-
on the hydrodynamic interference of the multi-rotor vertical tion design for reducing the axial force of a vaned mixed-
axis tidal current turbine. Renewable Energy, 169, 752–764. flow pump. Engineering Applications of Computational Fluid
https://doi.org/10.1016/j.renene.2021.01.055 Mechanics, 14(1), 882–896. https://doi.org/10.1080/199420
Sun, K., Ma, G., Wang, H., & Li, Z. (2019). Hydrodynamic per- 60.2020.1749933
formance of a vertical axis tidal current turbine with angular

You might also like