Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

1

EFFECT OF STAY VANES IN GRAVITATIONAL WATER VORTEX


POWER PLANT WITH SPIRAL BASIN

T.R. Bajracharya, A.B. Timilsina, R. Niraula, B. Rasaily, M. Lama, P. Pathak

Abstract: Demand for energy is ever so increasing and much of


the focus of energy research over the last decade has been
concentrated into developing an emission free and efficient
source of energy that can fulfil this growing demand.
Renewable energy such as micro and pico-hydropower remain
popular in developing countries for rural electrification and
considerably less cost of expenditures than large hydropower.
One such technology is Gravitational Water Vortex Power Plant
(GWVPP) which is capable of continuous energy generation by
converting the energy from a free surface water vortex to
rotational mechanical energy. This study aims to create a
reference for the design of a new and a compact type of basin
structure: a spiral-conical shaped basin, and also studies the
potential installation of stay vanes to influence the water vortex
parameter like tangential velocity which is a key parameter in
establishing turbine performance parameter. Three different
vane shapes: symmetrical, cambered and flat-bottomed were
used for runner design and the tangential velocity of these at
the plane of runner inlet was compared against the spiral-
conical basin without vanes. The analysis was first done
through model development in SOLIDWORKS, and then
respective simulation in ANSYS Fluent. Simulation results
showed that the spiral-conical basin model can maintain a base
efficiency of 52.59% and it was estimated that maximum
efficiency of around 80.32% can be achieved through proper
turbine runner shape. The basin models employed with flat-
bottomed shapes showed a possible utilization of vane shaped
structure and showed an increment of about 7% over the basin
without vanes. A 1kW pilot system with the same design
parameters can further validate the performance of the spiral-
conical GWVPP.

1 Introduction
Energy crisis remains one of the challenging issues facing the world with an
estimated 13 % of the world’s population still lacking an access to electricity. Due to
its lower cost and eco-friendliness, hydropower is seen as a viable solution to
mitigate the climate change issues and transition into clean, green and sustainable
energy. Small hydropower (SHP) less than 10 MW is particularly suited for rural
electrification due to its adaptability and positive impact to the development of the
rural communities. A report by World Small Hydropower Development Report, 2019
(WSHPDR) puts global installed capacity of SHP less than 10 MW at 78 GW
21st International Seminar on Hydropower Plants-Hydropower for future generations
09-11 November, Vienna, Austria
2

compared to the potential capacity of 229 GW [1]. One such micro-hydro-power


technology is the Gravitational Water Vortex Power Plant (GWVPP), which is an
ultra-low head turbine system which is capable of producing energy from low heads
ranging from 0.7m to 3m [3]. The GWVPP utilizes the energy from the rotating core of
a high angular velocity whirlpool which is generated by passing the water tangentially
in a vortex chamber. Since the first installation by Zotloterer at Obergrafendorf in
2006, a number of academic research on the turbine has been done and 19-22 full
scale installations are known to be operating mostly around Europe and Asia. The
average efficiency has been noted to be around 53 %, however, the GWVPP claims
superiority over its counterparts in terms of environmental friendliness, ease of
installation and cost effectiveness [7].

The subject of past researches on the GWVPP can be broadly categorized into two
categories: basin structure and the turbine runner. Mulligan, et. al. (2010) published
the first known experimental studies on the design and optimization of the GWVPP.
From the studies, the authors found an optimal range of orifice diameter to tank
diameter ratios (d/D) between 14% - 18% [2]. A study by Wanchat (2011) found that
a cylindrical tank with an orifice at the bottom centre with incoming flow guide was the
most suitable configuration to create the kinetic energy water vortex [3]. Dhakal, et.
al. (2014) studied the effect of dominant parameters for a conical shaped basin
structure. The study found that among the five different parameters: basin diameter,
notch angle, notch inlet width, cone angle and canal height, the inlet width (basin
opening) was found to be the most dominant and suggested that the opening be kept
small. For the basin diameter of 80 cm, inlet width of 10 cm, notch angle within 10° -
70°, cone angle of 23°, and canal height greater than 40 cm were suggested [4]. The
same year, Bajracharya, et. al. (2014) designed a conical shaped basin structure and
compared the performance with a previously installed cylindrical basin structure. The
tests concluded that conical shaped basin structure was preferred over and also
suggested runner with smaller number of blades [5]. A numerical and experimental
study by Timilsina, et. al. (2015) asserted that the maximum efficiency and power is
extracted at runner position of 65% - 75% from the top of the basin. The authors also
estimated the cost for a 7 kW GWVPP to be NRs. 922,232 for the metal parts [6]. A
detailed state of the art review performed by Timilsina, et. al. (2018) found that the
major large-scale installations of GWVPP had either a cylindrical or spiral typed flat
based basin structure. The authors acknowledged the performance of the conical
basin in which the flow of water becomes more axial towards the output thus
maximizing the tangential velocity difference and efficiency [7]. Dahal, et. al. (2019)
performed an experimental study of GWVPP with a booster runner installed in the
conical basin and concluded that an addition of booster runner near to the exit of the
basin corresponded to an increase in efficiency of 20.4% [8]. Bajracharya, et. al.
(2020) performed an analysis of geometrical parameters of runner design for the
GWVPP conical basin structure through numerical study and experiment. The study
recommended the runner height to basin height ratio of 0.31-0.32 [9]. A numerical
study on flat-spiral shaped basin structure was conducted by Huwae, et. al. (2020)
which compared flow streamlines between a flat-spiral basin and a conical basin and
concluded that spiral circulation tank is better than the cylindrical tank for same size
generation system [10].

21st International Seminar on Hydropower Plants-Hydropower for future generations


09-11 November, Vienna, Austria
3

From the past studies, the trend for the most optimal basin structure can be seen
going from a simple flat cylindrical shape to a cylindrical-conical shaped to a flat
spiral shape, with each basin structure showing an increase in water vortex
parameters and thus, efficiency over the preceding ones. This study concerns on the
study of a new type of basin structure for GWVPP: a spiral-conical shaped basin
structure. Suitable design parameters for a 100 cm base spiral diameter were
designed with reference to the findings from the past studies. This basin design was
then employed with three different stay vanes shapes: symmetrical ones used in
other reaction turbines such as Francis, cambered shaped and a flat-bottomed shape
to investigate the effect of the vanes on the flow of water vortex. Computation Fluid
Dynamics (CFD) simulations were then performed thoroughly on ANSYS Fluent.
Comparisons of tangential velocity between the different models at a suitable runner
inlet plane was then performed and discussed.
2 CFD Model Development
Computational Fluid Dynamics (CFD) is used to simulate the complexity vortex flow
in the basin of GWVPP [10]. CFD replaces the complex governing equations of fluids
into discretized algebraic forms which are easier to solve [9]. All CFD processes
generally include three main elements: (i) a pre-processor, (ii) solver, and post
processor. The air-core vortex is assumed to be steady, axisymmetric and
incompressible flow for which the continuity equation and the Navier-Stokes
equations in cylindrical coordinates are described as follows [5].

∂𝑉𝑟 ∂𝑉𝑧 𝑉𝑟
+ + =0
∂r ∂z r
∂𝑉𝜃 ∂𝑉𝜃 𝑉𝑟 𝑉𝜃 ∂2 𝑉𝜃 ∂𝑉𝜃 𝑉𝜃 ∂2 𝑉𝜃
𝑉𝑟 + 𝑉𝑧 − = 𝑣( 2 + − + )
∂r ∂z r ∂𝑟 𝑟 ∂𝑟 𝑟 2 ∂𝑧 2
∂𝑉𝑟 ∂𝑉𝑟 𝑉𝜃2 ∂𝜌 ∂2 𝑉𝑟 ∂𝑉𝑟 𝑉𝑟 ∂2 𝑉𝑟
𝑉𝑟 + 𝑉𝑧 − + = 𝑣( 2 + − + )
∂r ∂z r 𝜌 ∂𝑟 ∂𝑟 𝑟 ∂𝑟 𝑟 2 ∂𝑧 2
∂𝑉𝑧 ∂𝑉𝑧 ∂𝜌 ∂2 𝑉𝑧 ∂𝑉𝑧 ∂2 𝑉𝑟
𝑉𝑟 + 𝑉𝑧 + = 𝑔+𝑣( 2 + + )
∂r ∂z 𝜌 ∂𝑧 ∂𝑟 𝑟 ∂𝑟 ∂𝑧 2

Where, vϴ is the tangential velocity, vr is the radial velocity and vz is the axial velocity.
Likewise, ρ is the density, 𝑣 is the kinematic viscosity and g is the gravity.

Fluid domains were designed in SOLIDWORKS and respective simulations were


performed on ANSYS Fluent 19.2. After each simulation, analysis was done and
necessary changes were made on succeeding models.

2.1 Basin Design Parameters


The design parameters for the spiral-conical basin relates closely to that of basin
design from conical basin. One key difference between the spiral-conical basin and
its counterparts lies in the design of the cone. On the conical basin, the cone
diameter is concentric and equal to that of the cylinder diameter, whereas, on the
spiral-conical basin, the cone diameter is concentric with the base circle from which
the spiral shape is created. The cone diameter is not fixed on this model but is
constrained by the cone angle and the gross head available.
21st International Seminar on Hydropower Plants-Hydropower for future generations
09-11 November, Vienna, Austria
4

Table 1: Design parameters for the basin structure


Parameters Dimensions (cm)
Spiral basin diameter (D) 100
Total height of the basin (H) 101.23
Canal height 60
Inlet notch width 15
Notch angle 13°
Cone angle 23°
Cone diameter (Dc) 50
Outlet diameter (d) 15

Since the cone angle is kept constant, as the cone diameter is increased, the cone
height increases thus increasing the total height of the basin. For the study, the cone
diameter was set as 50 cm which corresponds to the total basin height of 101.23 cm.
A similar setup of a conical basin of height 1.4 m has already been installed at the
Center for Energy Studies (CES), IOE, Nepal, thus the proposed spiral-conical model
can be fabricated and installed at the facility.

Figure 1: Spiral-conical basin design parameters

2.2 Vanes design methodology


For the basin model with stay vanes, vanes of different shapes and sizes were
employed on the periphery of the cone diameter of the spiral-conical basin. The basin
for this study will have a reaction type turbine runner for the experimental setup. The
design methodologies for stay vanes were taken from reaction turbine such as that of
Francis turbines. A brief literature review on reaction turbine suggests that the design
procedure varies for different reaction turbines. Desai, et. al. (2010) performed a
validation study of stay vanes in Francis turbine using CFD. The authors
recommended using conventional stay vanes shape which has ease of
manufacturing and the authors employed an aerofoil shaped NACA 2520-63 for their
investigation [11]. Relevant literature on design methodology on stay vanes were not
21st International Seminar on Hydropower Plants-Hydropower for future generations
09-11 November, Vienna, Austria
5

sufficient hence some analogies were also taken from guide vanes design. The
literature review suggested that the stay vanes design is an iterative design process
whereas the guide vanes design depends upon parameters such as shape, size and
orientation.

The purpose of this study was to see whether a suitable stay vane arrangement can
be employed in GWVPP without disturbing the free vortex flow and to investigate the
effect of these vanes on the water vortex parameters. In essence, the vanes in
GWVPP will act both as a stay vanes as well as guide vanes of Francis turbine.
Thapa, et. al. (2016) used CFD to design and develop a guide vane for a low-speed
number Francis turbine. The speed number (ω) is a speed ratio parameter which is
used to configure guide vanes design. Assuming the proposed basin design of the
study, for flow rate of 0.165 m3/s and rotational speed of 95 rev/min, the speed
number calculation for the GWVPP comes out to be 0.4334. For their study, the
authors took Jhimruk Hydroelectric Centre (JHC) as a case study, which is a low-
speed number Francis turbine of 0.32. The hydropower employs a Francis turbine
with a NACA 0012 profile guide vanes, which was chosen as the starting premise for
the design of stay vanes for this study [12].

Typically, four different parameters, chord length (c.l.), number of vanes (z), vane
shape and vane outlet angle can be considered for the design of vanes. The effect of
three different vane shapes: symmetrical, cambered and flat-bottomed vanes were
investigated in the study and the overall design procedure for the vanes was an
iterative approach with each simulation on the respective vanes providing essential
feedbacks for the design parameters of the successive models. The following design
considerations were taken for the design of vanes model:

i. For some of the model, the chord length and the number of vanes relation was
kept as:
𝜋. 𝐷𝑐
0.7 𝑐. 𝑙. =
𝑧

ii. The orientation of vanes was determined by the chord line and lower camber,
either of which was oriented tangential to the periphery of cone. This
orientation would ensure that the vortex flow isn’t drastically disturbed.

Figure 2: Orientation of vanes in the basin

21st International Seminar on Hydropower Plants-Hydropower for future generations


09-11 November, Vienna, Austria
6

iii. Apart from NACA 0012, other aerofoil shapes were chosen such that their
database (the coordinates of the lower camber and upper camber) were
readily available and the respective vanes would conform to the require
tangential orientation.

Figure 3: From top-left to bottom right: NACA 0012, NACA 2305, NACA 6405 and S7055 (10.5%)
vanes shapes

The stay vanes models will be named as S.M. followed by a number which indicates
the linearity of the simulation of the models in an ascending order.

Table 2: Specifications of different stay vanes GWVPP


models
S.N. Vane shape Chord No. of
length blades
(cm) (z)
S.M.1 NACA 0012 13.2 17
S.M.2 NACA 0012 18.7 12
S.M.3 NACA 2305 13.2 17
S.M.4 NACA 2305 13.2 9
S.M.5 NACA 2305 22.4 10
S.M.6 NACA 6405 13.2 10
S.M.7 NACA 6405 13.2 10
S.M.8 S7055 (10.5%) 13.2 17
S.M.9 S7055 (10.5%) 12 17
S.M.10 S7055 (10.5%) 12 8
S.M.11 S7055 (10.5%) 12 24
S.M.9.1 S7055 (10.5%) 12 17
S.M.10.1 S7055 (10.5%) 12 8
Note: Stay vanes models from S.M.1 to S.M.11 have vanes
height of 10 cm. The models S.M.9.1 and S.M.10.1 have
vanes height of 15 cm.

2.3 CFD Mesh


The fluid domains were then imported into ANSYS Fluent where subsequent
simulations were performed. Initially, the default mesh was generated, which was
refined locally to fulfil the criteria for an acceptable mesh condition. The whole
geometry was separated into sweepable and non-sweepable domain. Hexa-elements

21st International Seminar on Hydropower Plants-Hydropower for future generations


09-11 November, Vienna, Austria
7

were used in sweepable domain, whereas, for non-sweepable domain, the presence
of vanes and the volute shape of the basin resulted in complexity, therefore the
meshing was performed using tetrahedral elements.

Table 2: Mesh quality obtained on average for all models


Criteria Mesh quality obtained
Aspect ratio max < 20
Skewness 0.8 – 0.87
Minimum orthogonal quality 0.11 – 0.2
Smoothness < 15%

The simulation has been performed for a steady state analysis by solving 3D
Reynolds averaged Navier-Stokes equations (RANS) with a control volume-based
technique and standard k-ω SST turbulence model. The k-ω Shear Stress Transport
model switches to k-epsilon behaviour in the free stream and avoids the k-omega
problem of being sensitive to the inlet free stream turbulence properties while also
behaving as the standard k-ω model which is best used for near-wall treatment.
Semi-Implicit Method for Pressure Linked equations (SIMPLE) velocity-pressure
coupling algorithm was employed along with the second order pressure interpolation
and second order discretization scheme for turbulence and momentum.

Figure 4: Boundary conditions

The simulation was performed with the inlet as velocity inlet with the water flow set as
0.8 m/s. No-slip conditions was set at the wall and the outlet was characterized as
pressure outlet. The upper surface of the domain was set to symmetry.

3 Results and Discussions


For a reaction-based turbine runner, the efficiency of the turbine can be calculated
as:

(𝑣𝛳1 − 𝑣𝛳2 ) ∗ 𝑢
𝜂=
𝑔𝐻

Where, u represents the turbine runner speed at inlet plane and, v ϴ1 and vϴ2
represents the tangential velocity at the turbine runner inlet and outlet respectively.

21st International Seminar on Hydropower Plants-Hydropower for future generations


09-11 November, Vienna, Austria
8

The plot of the tangential velocity at the plane of the runner inlet gives us the required
value of vϴ1. The value of vϴ2 depends upon the turbine runner. From the equation, it
is evident that maximizing the value of (vϴ1 - vϴ2) increases the turbine efficiency. For
the designed GWVPP model, the turbine runner inlet plane was calculated as 0.65 of
the total height which was 66 cm from the top of the basin as per the optimal position
recommended from the previous studies. The tangential velocity graphs for all the
models were then plotted and compared at the diameter of this circular plane.

First, the tangential velocity at the plane of interest for the spiral-conical model and
the basin model employed with symmetrical stay vanes similar to that of Francis
turbine was plotted. The graph showed that the tangential velocity decreased for the
symmetrical vanes model. This could be attributed to the size of the symmetrical
vane which is noticeably thicker than other vanes for the same chord length.

Figure 5: Tangential velocity graph of spiral-conical and symmetrical vanes model

The vortex formed in GWVPP is characterized as free-surface vortex (though,


external energy is continuously supplied by the incoming water). For the vanes to not
have diminishing effect on tangential velocity, the water flow must be properly guided
at the inlet of the vanes without any significant presence of losses such as reverse
flow loss. A cambered shaped profile would mean more guided flow at the inlet of the
vane so, the next models were employed with NACA 2305, and NACA 6405 vane
shapes. No such substantial improvement was found from the NACA 2305 model
over the spiral-conical model, however, graphical plots of S.M. 5 and S.M. 6
suggested that for the same number of vanes, the chord length calculations seemed
irrelevant and the shorter chord length minimizes the loss of water.

21st International Seminar on Hydropower Plants-Hydropower for future generations


09-11 November, Vienna, Austria
9

Figure 6: Tangential velocity graph of spiral-conical and 2% cambered vanes models

The 6% cambered vanes profile did show an increase in tangential velocity over the
symmetrical and the 2% cambered vane shapes. However, the increment wasn’t
discernible enough to justify its installation over the spiral-conical model.

Figure 7: Tangential velocity graph of spiral-conical and 6% cambered models

For the last stay vanes shape, a flat-bottomed aerofoil shape S7055 (10.5%) was
chosen. The spiral models S.M.9 and S.M.10 showed a broader tangential velocity
graph as compared to the base spiral-conical model. The plots for the vanes model
weren’t however symmetrical about the centre of the cone but rather skewed towards
the plane from which the inlet water flow is initialized. The S.M.9 and S.M.10 models
were modified by increasing the vanes height to 15 cm. The tangential velocity plots
for the S.M.9.1 and S.M.10.1 was observed to be slightly broader over the diameter
as compared to earlier counterparts.

21st International Seminar on Hydropower Plants-Hydropower for future generations


09-11 November, Vienna, Austria
10

Figure 8: Tangential velocity graph of spiral-conical and flat-bottomed vanes models

Figure 9: Tangential velocity graph of spiral-conical and flat-bottomed vanes (15 cm height) models

The radius of the cone at the turbine runner inlet position is 22.45 cm. Assuming a
constant clearance of 4 cm, the outer edge of the turbine runner at inlet will be at a
distance of 18.45 cm from the centre of the cone. The values of tangential velocity at
this point were extrapolated for the spiral-conical models & the models that showed a
potential increment.

21st International Seminar on Hydropower Plants-Hydropower for future generations


09-11 November, Vienna, Austria
11

Table 3: Values of tangential velocity at the inlet plane


Model Tangential velocity(m/s) at 18.45 cm
from the centre of the cone
Spiral- 5.84427
conical
S.M.9 6.25695
S.M.9.1 6.06460
S.M.10.1 6.11560

The flow of water in the GWVPP is a swirling motion where the tangential component
of velocity is more dominant in the upper region. Since the water exits the basin
axially at the outlet, towards the cone, the water flow becomes more axial. The
tangential velocity at the outlet can be minimized by increasing the exit angle of the
turbine blade at outlet.

The efficiencies for a range of values of tangential velocity at the turbine runner outlet
was calculated for the turbine runner inlet at 18.45 cm from the centre of the cone for
the following value of other parameters:

u = ω*r = 9.94838 * 0.1845 = 1.83548 m/s


g = 9.81 m/s2
H = 1.0123 m
Theoretical power input = ρ*g*Q*H = 1000*9.81*0.168*1.0123 = 1668.35 W

Table 4: Efficiency for different values of tangential velocity at


outlet
Model Efficiency when vϴ2 =
1.5 m/s 2 m/s 3 m/s
Spiral-conical 80.32 % 71.07 % 52.59 %
S.M.9 87.95 % 78.70 % 60.22 %

From the above table, it can be seen that even a slight increment in the tangential
velocity at inlet leads to a significant increment in turbine efficiency. All of the stay
vanes model implemented with a flat-bottomed shaped vanes showed an increment
in tangential velocity over the spiral-conical basin without any vanes. This also
suggests that conventional vanes design methodology used in other reaction turbines
isn’t exactly analogous for a possible vane installation in GWVPP.

For the same inlet basin height and basin diameter, the conical basin takes up more
height than for the spiral-conical basin due to the design parameter of the cone. From
the above table, it is evident that even when the flow at the turbine runner outlet is
considered to be significantly tangential (which isn’t normally the case since all of the
water exits axially), the spiral-conical model still shows an impressive efficiency of
52.59%. For the real-life installation purpose, loss of about 10% can be nominally
expected but this number is still impressive taking into the consideration the
compactness of the spiral-conical basin structure and greater outlet tangential
velocity.

A more accurate multiphase simulation on GWVPP including a suitable reaction type


runner can be performed to validate the efficiency of the spiral-conical basin and the

21st International Seminar on Hydropower Plants-Hydropower for future generations


09-11 November, Vienna, Austria
12

possible implementation of vanes. The simulation can also give us a possible range
of tangential velocity at corresponding runner outlet which can then be taken as a
reference for future simulations on GWVPP with other basin structures.

4 Conclusions
This study has analysed water vortex parameters of a new spiral-conical shaped
basin structure of GWVPP and the potential installation of stay vanes in the basin.
The different basin models with stay vanes were analysed and compared in terms of
tangential velocity at the inlet plane of turbine runner. CFD analysis showed that the
spiral-conical basin without any vanes can maintain an efficiency of more than 52.59
%, though this value can be estimated to be much higher as the tangential velocity at
the runner outlet is considerably axial. The simulations also showed that when a flat-
bottomed shaped vanes was installed around the periphery of the cone, there is an
increment in tangential velocity, without any significant distortion to the free surface
water vortex formed in the basin. This increment, while minute, increases the
efficiency of the turbine by a significant margin, which suggests that the tangential
velocity is a very sensitive parameter. The maximum efficiency of the turbine with and
without the flat-bottomed vanes were noted as 80.32% and 87.95% respectively,
though a decrease of around 10% can be expected for real life installation. While the
past studies on the basin characterized the outlet velocity as a key performance
parameter, this study claims that the tangential velocity at the outlet should be taken
as a basis for GWVPP performance as the turbine runner for the spiral-conical basin
becomes primarily reaction type turbine.

A pilot installation of 1kW spiral-conical basin GWVPP can give the necessary turbine
performance parameters and can further validate the results of this research. The
proposed spiral-conical basin in this research should cost less than the 1 kW conical
basin GWVPP of same size due to the compactness of the spiral-conical basin.

5 References

[1] Hydropower Status Report, International Hydropower Association, 2022.

[2] S. Mulligan and P. Hull, Design and Optimisation of a Water Vortex Power
Plant, Institute of Technology, Siligo, 2010.

[3] S. Wanchat, Preliminary Design of a Vortex Pool for Electrical Generation,


Journal of Computational and Theoretical Nanoscience, 2011.

[4] S. Dhakal, A.B. Timilsina, R. Dhakal and D. Fuyal, Effect of Dominant


Parameters for Conical Basin: Gravitational Water Vortex Power Plant, IOE
Graduate Conference (2014), 380-386.

[5] S. Dhakal, S. Nakarmi, P. Pun, A.B. Thapa and T.R. Bajracharya,


Development and Testing of Runner and Conical Basin for Gravitational Water
Vortex Power Plant, Journal of the Institute of Engineering (2014), 140-148.

21st International Seminar on Hydropower Plants-Hydropower for future generations


09-11 November, Vienna, Austria
13

[6] S. Dhakal, A.B. Timilsina, R. Dhakal, D. Fuyal, T.R. Bajracharya, H.P. Pandit,
N. Amatya and A.M. Nakarmi, Comparison of cylindrical and conical basins
with optimum position of runner: Gravitational water vortex power plant,
Renewable and Sustainable Energy Reviews 48 (2015), 662-669.

[7] A.B. Timilsina, S. Mulligan and T.B. Bajracharya, Water vortex hydropower
technology: a state-of-the-art review of development trends, Springer Nature,
2018.

[8] N. Dahal, R.K. Shrestha, S. Sherchan, S. Milapati, S.R. Shakya and A.K. Jha,
Performance Analysis of Booster based Gravitational Water Vortex Power
Plant, Journal of the Institute of Engineering 15 (2019), 90-96.

[9] T.R. Bajracharya, S.R. Shakya, A.B. Timilsina, J. Dhakal, S. Neupane, A.


Gautam and A. Sapkota, Effects of Geometrical Parameters in Gravitational
Water Vortex Turbines with Conical Basin, Journal of Renewable Energy,
Hindawi, 2020.
[10] R, Huwae, A. Susatyo, R.A. Subekti, H, Sudibyo and D.S. Khaerudini, A
Parametric of Basin Geometry of Gravitational Water Vortex Power Plant,
International Conference on Sustainable Energy Engineering and Application,
2020.
[11] J, Desai, A, Roghella, V.A. Soni and V. Chauhan, Validation of
Hydraulic Design of Spiral Casing and Stay Vanes of Francis Turbine Using
CFD, 37th International & 4th National Conference on Fluid Mechanics and
Fluid Power, 2010.
[12] B.S. Thapa, C. Trivedi and O.G. Dahlhaug, Design and development of
guide vane cascade for a low-speed number Francis turbine, Journal of
Hydrodynamics, 2016.

21st International Seminar on Hydropower Plants-Hydropower for future generations


09-11 November, Vienna, Austria

You might also like