You are on page 1of 12

Effective Basin–Blade Configurations of a Gravitational

Water Vortex Turbine for Microhydropower Generation


Nauman Hanif Khan 1; Taqi Ahmad Cheema 2; Javed Ahmad Chattha 3;
and Cheol Woo Park 4

Abstract: Among various microhydropower plants, gravitational water vortex power plants are emerging because of their simple installa-
Downloaded from ascelibrary.org by Universitas Indonesia on 01/22/19. Copyright ASCE. For personal use only; all rights reserved.

tion, reduced setup time, and minimal maintenance cost. The methodology and the selection of a suitable basin and turbine blade combi-
nations are yet to be explained by researchers. This study attempts to investigate the parameters that affect the formation and strength of a
vortex for efficient power generation using an artificially induced air-core vortex. Computational fluid dynamics based on a two-phase flow
analysis of the vortex formation at different basin parameters resulted in the selection of a suitable configuration of the basin. The basin was
then used in the blade analysis with different blade shapes at various load conditions. An experimental setup was also fabricated for the
validation of the numerical results, in which a close agreement was observed. The proposed methodology could be used to determine the plant
specifications and the blade size and shape for various flow rates and heads. The role of vortex height in determining the performance of the
gravitational water vortex turbine was explored. Among the four types of blades used in the study, cross-flow blades have shown the best
efficiency for the same discharge and head conditions. DOI: 10.1061/(ASCE)EY.1943-7897.0000558. © 2018 American Society of Civil
Engineers.
Author keywords: Water vortex; Turbine; Basin; Air core; Vortex height; Blade; Efficiency.

Introduction Despite having the previously mentioned advantages over other


types of conventional turbines, limited work has been conducted
Free-surface vortices are formed at the intake of hydraulic struc- on the research and development of GWVTs. Since the completion
tures, such as spillways and pump sumps, mainly because of the of the first ever GWVT power plant in 2004, researchers have
Coriolis force. The energy in a vortex can be harnessed by intro- paid some attention mainly to the optimization of the basin and
ducing water flow tangentially into a cylindrical basin with an or- blade. For example, Singh and Nestmann (2009, 2011) completely
ifice at the bottom. The combined effect of the localized low modified the GWVT layout by using guide vanes as the stator and
pressure at the orifice and the circulation that is induced as a result the blades as the rotor. Their blade optimization methodology in-
in the tangential entry through inlet causes the water to restructure cluded the increase of efficiency by reducing the discharge through
into a vortexlike flow pattern (Venukumar 2013). The gravitational the turbine. Similarly, Venukumar (2013) experimentally demon-
potential energy of the entering water is entirely converted into strated that the power produced by GWVT is greater than any
kinetic energy, which can be extracted from the water by using other conventional hydropower plants for the same conditions.
a coaxial turbine with a vertical axis. This turbine is known as Recently, Power et al. (2015) performed a detailed experimental
the gravitational water vortex turbine (GWVT). GWVT is a suitable investigation of the effects of varying flow rates, inlet conditions,
choice as a hydraulic energy converter because it is safer for aquatic and size and number of blades on turbine speed, torque, and vortex
life, the environment, and water quality and because of its lower height.
speed operation, easier installation, no reservoir requirement, lower With the advancement of computational tools in the last
operating and manufacturing costs, and higher power output decade, the computational fluid dynamics (CFD) approach has be-
compared with other types of microhydro turbines for the same come an effective alternative in simultaneously investigating the
conditions (Khan et al. 2008; Dhakal et al. 2014; Mohanan 2016; various factors that affect the physics of a phenomenon. Several
Elbatran et al. 2015). researchers have also applied the same approach to the concept
of GWVT. For example, Wanchat and Suntivarakorn (2012) and
1 Wanchat et al. (2013) used an experimentally validated CFD
Graduate Student, Dept. of Mechanical Engineering, GIK Institute of
Engineering Sciences and Technology, Topi, KPK 23460, Pakistan. model to simulate the vortices formed in three different types
2
Assistant Professor, Dept. of Mechanical Engineering, GIK Institute of of basin geometries. In another study, Shabara et al. (2015) used
Engineering Sciences and Technology, Topi, KPK 23460, Pakistan. numerical simulation to investigate the optimum vortex pool con-
3
Professor, Dept. of Mechanical Engineering, GIK Institute of figuration of a GWVT. The efficiency of the setup was found to be
Engineering Sciences and Technology, Topi, KPK 23460, Pakistan. approximately 40% between 28 and 38 rpm. Similarly, Dhakal
4
Professor, School of Mechanical Engineering, Kyungpook National et al. (2015) numerically analyzed different basin geometries for
Univ., 80 Daehak-ro, Buk-gu, Daegu 41566, South Korea (corresponding vortex formation and velocity measurement. They proposed that
author). Email: chwoopark@knu.ac.kr
the optimum position of runner in the basin is 65% from the
Note. This manuscript was submitted on October 11, 2017; approved on
February 12, 2018; published online on May 19, 2018. Discussion period top of vortex because this is the point at which the maximum
open until October 19, 2018; separate discussions must be submitted for velocities can be achieved.
individual papers. This paper is part of the Journal of Energy Engineering, Vortex formation has been explained fairly by several research-
© ASCE, ISSN 0733-9402. ers. However, studies on the parameters that affect the vortex

© ASCE 04018042-1 J. Energy Eng.

J. Energy Eng., 2018, 144(4): 04018042


formation in basins are scarce. The data on the selection and design Materials and Methods
of GWVT basin based on in-flow conditions, which will ensure the
formation of a strong and stable vortex, are unavailable because of The methodology adopted for the CFD analysis may be divided
the negligible work that has been performed on the aspect ratios of into two steps. The first step is the selection of a basin with a full
the basin and width and depth of the inlet. Moreover, numerous air-core vortex, high gain in tangential velocities, and reasonable
authors simulated different blade profiles to extract energy using vortex height. The second step is the comparison of four blade
GWVT. However, their numerical results were not supported by shapes to achieve maximum output power. This comparison yields
the experimental validation. Most importantly, the previous authors the optimum blade shape to use in GWVT.
relied on the single-phase flow of water for their analysis, whereas
the actual physics of the water vortex energy conversion demands Selection of Basin
the presence of air to be considered in a GWVT. The role of vortex The literature review suggests that the increase in the mean tangen-
height is an important parameter in the physics of vortex formation tial velocity of water in a vortex basin makes the water motion tan-
to determine the performance of a GWVT that is yet to be explored gential instead of axial because of the increase in vortex height
by researchers.
Downloaded from ascelibrary.org by Universitas Indonesia on 01/22/19. Copyright ASCE. For personal use only; all rights reserved.

(Kueh et al. 2014; Wanchat et al. 2013; Odgaard 1986). Moreover,


The aforementioned concerns have motivated the authors to Wanchat and Suntivarakorn (2012) showed that a cylindrical tank
conduct a detailed study on the selection of the most suitable basin with a discharge hole at the bottom center is the most suitable con-
and blade configuration of a GWVT to achieve maximum power figuration to generate a water vortex along with a tangential water
output and efficiency for given flow conditions. This study was per- entry into a basin. Moreover, the most feasible outlet to basin diam-
formed in two main parts. In the first phase, the computational ap- eter ratio is 0.14–0.18. In addition, the formation of a fully devel-
proach was implemented for the analysis of the factors that affect oped air core increases the potential for power production from the
the vortex formation, the selection of the most suitable basin con- vortex. Therefore, selecting a basin that generates a free surface and
figuration, and the comparative analysis of the various blade pro- fully developed air-core vortex with a maximum possible height
files under two-phase flow conditions. In the second phase, the and high velocities at the vortex core has been the main focuses
results obtained by the numerical simulations were validated using of this study. Thus, a reference basin was designed, and its dimen-
the experiments for the same flow conditions. The in-line numerical sions are shown in Fig. 1(a).
and experimental results are expected to help improve turbine To increase the vortex strength, the vortex dependency param-
efficiency. eters of the basin geometry were individually varied from the

Fig. 1. (a) Reference basin model; and (b) blade profiles used in the study.

© ASCE 04018042-2 J. Energy Eng.

J. Energy Eng., 2018, 144(4): 04018042


Table 1. Sets of values used for the basin analysis
Serial Basin diameter Outlet diameter to Basin height Aspect ratio Inlet width Inlet depth Flow velocity Mass flow
number (mm) basin diameter ratio (mm) (D=H) (mm) (mm) (m=s) (kg=s)
1 500 0.14 500 1 150 150 0.1 2.5
2 500 0.14 500 1 150 150 0.2 5.0
3 500 0.14 500 1 150 150 0.3 7.5
4 500 0.14 500 1 150 150 0.4 10.0
5 500 0.14 500 1 150 150 0.6 12.5
6 500 0.14 500 1 150 50 0.3 2.5
7 500 0.14 500 1 150 100 0.3 5.0
8 500 0.14 500 1 150 150 0.3 7.5
9 500 0.14 500 1 150 200 0.3 10.0
10 500 0.14 500 1 150 250 0.3 12.5
11 500 0.14 500 1 50 150 0.3 2.5
Downloaded from ascelibrary.org by Universitas Indonesia on 01/22/19. Copyright ASCE. For personal use only; all rights reserved.

12 500 0.14 500 1 100 150 0.3 5.0


13 500 0.14 500 1 150 150 0.3 7.5
14 500 0.14 500 1 200 150 0.3 10.0
15 500 0.14 500 1 250 150 0.3 12.5
16 500 0.14 250 0.5 150 150 0.3 7.5
17 500 0.14 375 0.75 150 150 0.3 7.5
18 500 0.14 500 1 150 150 0.3 7.5
19 500 0.14 625 1.25 150 150 0.3 7.5
20 500 0.14 750 1.5 150 150 0.3 7.5
21 500 0.13 500 1 150 150 0.3 7.5
22 500 0.14 500 1 150 150 0.3 7.5
23 500 0.15 500 1 150 150 0.3 7.5
24 500 0.16 500 1 150 150 0.3 7.5
25 500 0.17 500 1 150 150 0.3 7.5
26 400 0.20 500 0.8 150 150 0.3 7.5
27 500 0.16 500 1 150 150 0.3 7.5
28 600 0.13 500 1.2 150 150 0.3 7.5
29 700 0.11 500 1.4 150 150 0.3 7.5
30 800 0.10 500 1.6 150 150 0.3 7.5

reference values, while all others were maintained constant and an- different authors (Li et al. 2008; Venukumar 2013; Dhakal et al.
alyzed for the tangential velocities of the water in the vortex core, 2015). The shapes include inverted conical blades (Profile 1),
vortex height, and the quality of air core ranging from 0 (for no air cross-flow blades (Profile 2), curved rectangular blades (Profile 3),
core) to 1 (for full air core). Aside from varying the geometric and twisted blades (Profile 4), as shown in Fig. 1(b). The dimen-
parameters, the flow conditions were also varied to analyze the effect sions of the selected blades are completely dependent on the vortex
of flow velocities on vortex formation. Parameters including basin formed in a particular basin. The vortex profile was analyzed, re-
diameter (0.4–0.8 m), outlet to basin diameter ratio (0.13–0.17), sulting in the designing of the blades. In a vortex, the layers of fluid
basin height to diameter ratio (0.5–1.5), inlet channel width ratio flow were spiral and had tangential velocities greater than the radial
(0.1–0.5), inlet channel depth ratio (0.1–0.5), and inlet velocity and axial velocities with the optimum position for the blade instal-
(0.1–0.6 m=s) were analyzed to investigate their effect on vortex lation at 35% of the vortex height from the bottom of the vortex
and tangential velocities. Table 1 shows a summary of the cases that (Dhakal et al. 2014). In view of the vortex profile of each blade,
were analyzed in this study. The obtained results enabled a basin to the outer diameter of the blades was selected such that it should not
generate a full air-core vortex, along with a considerable increase in be less than the minimum distance between the water surfaces on
the kinetic head using the aforementioned methodology (Table 2). both sides of the blade, at a position of 35% from the bottom of the
The results of the basin analysis and selection of the most suitable basin. This arrangement enabled the water to transfer maximum
configuration have already been reported by Chattha et al. (2017). kinetic energy to the blades for power generation. This arrangement
caused complexity when the vortex height decreased with the an-
Selection of Blades gular velocity of the rotating blades.
The basin was divided into two domains to enable rotation of the
Four blade profiles were analyzed for the power production and blades in the basin. An outer stationary domain was referred to as
vortex distortion in the basin of a GWVT using ANSYS CFX. the basin domain, and a central cylindrical rotating domain, which
The selection of the blade profiles was based on the research of contained the blades, was referred to as the blade domain. The po-
sition of the blade walls inside the cylindrical domain was selected
Table 2. Optimized values of basin parameters
to maintain the height of the blade at 100 mm above the bottom of
the basin and to ensure the submergence of the blade in the vortex at
Basin parameter Optimized value (mm) all vortex heights.
Basin diameter 500
Basin height 500
Outlet diameter 80 Mathematical Modeling and Governing Equations
Inlet width 100 The free-surface air-core vortex was formed as a result of
Inlet depth 100
the interface between air and water (Li et al. 2008;

© ASCE 04018042-3 J. Energy Eng.

J. Energy Eng., 2018, 144(4): 04018042


Wanchat and Suntivarakorn 2012). Therefore, the analysis was required for the vortex and the turbine because air was present in-
based on the multiphase Eulerian fluid approach, in which two side the runner in the regions that were not submerged in water.
fluids, air and water, occupy the same domain and velocity fields.
The temperature of the domain was set to 25°C, and the reference
pressure was set to 101.325 kPa (1 atm). The buoyancy reference Numerical Method and Implementation
density was set to the density of air at 25°C, which is equal
to 1.185 kg=m3 . The shear stress transport (SST) model was developed to overcome
The governing equations for a steady, incompressible, viscous, the deficiencies in the k–ω and baseline k–ω models. The use of the
and turbulent vortex flow were the continuity and the Navier- k-ω formulation made the turbulence model usable for the inner
Stokes equations. These equations are described in cylindrical parts of the boundary layer, as well as for near-wall regions (Ansys
coordinates as follows (Wang et al. 2010): 2015). Therefore, Shabara et al. (2015) preferred the use of the SST
model over other models. They used the SST model to simulate the
∂V r ∂V z V r artificial vortex for a microhydropower plant using the multiphase
þ þ ¼0 ð1Þ
∂r ∂z r Eulerian fluid approach and performed experimental validation for
Downloaded from ascelibrary.org by Universitas Indonesia on 01/22/19. Copyright ASCE. For personal use only; all rights reserved.

the obtained results. The ANSYS built-in feature was utilized to


 2 
∂V θ ∂V θ V r V θ ∂ V θ ∂V θ V θ ∂ 2 V θ optimize a mesh structure automatically by meshing the blades us-
Vr þ Vz − ¼ν þ − þ ð2Þ ing automatic tetrahedral mesh elements in ANSYS ICEM. This
∂r ∂z r ∂r2 r∂r r2 ∂z2
process would create finite smooth transition blocks with high re-
 2  finement on walls to obtain the final mesh. The mesh sensitivity
∂V r ∂V r V 2θ ∂p ∂ V r ∂V r V r ∂ 2 V r analysis of the basin domain was performed based on the mass flow
Vr þ Vz − þ ¼ν þ − þ
∂r ∂z r ρ∂r ∂r2 r∂r r2 ∂z2 and velocity of water, while mesh sensitivity analysis of the blade
ð3Þ domain was performed based on the torques on the blade surfaces
(Chen et al. 2012). Results of mesh sensitivity analysis can be
 2  found in the Appendix. Three different numbers of mesh elements
∂V z ∂V ∂p ∂ V z ∂V z ∂ 2 V r
Vr þ Vz z þ ¼gþν þ þ ð4Þ were used for both basin and blade analysis by calculating the vor-
∂r ∂z ρ∂z ∂r2 r∂r ∂z2 tex height [Fig. 10(a)] and the torque on blades [Fig. 10(b)], respec-
tively. For basin analysis, the inlet velocity of water was varied,
where V r , V θ , and V z = radial, tangential, and axial velocities, re- whereas 110 rpm was used for computation of torque using four
spectively g = gravitational acceleration; ν = kinematic viscosity; different blade profiles. The number of tetrahedral elements used
and ρ = fluid density. for the basin and blade domains were 4,816,342 and 8,922,533,
The following expressions were used for computing various respectively. The element size in both domains was 0.001 with
parameters in the turbine analysis. If N is the speed of the turbine a refinement of five layers at the blade surface.
in revolutions per minute, then the angular velocity (input param- The present study used commercial code ANSYS CFX, which
eter) can be calculated as discretizes the governing equations using a finite-volume approach
based on elements with cell vertex formulation. The common
Angular velocityðωÞ ¼ 2 × π × N=60 ð5Þ
solver parameters used include an advection scheme and turbulence
numerics. The selected advection scheme was high resolution,
If the torque on blade assembly is T, then power can be calcu- whereas the first-order upwind scheme was selected from turbu-
lated as follows:
lence numerics. The convergence control and its criterion were
Brake horse power ðBHPÞ ¼ Output power ¼ T × ω ð6Þ set different for the two domains. In the initialization of the do-
mains, the basin domain was identified as the stationary domain,
whereas the blade domain was set to rotate with increments of
Input power based on total head ðIHPÞ ¼ γHQ ð7Þ 25 rpm beginning from 50 up to 200 rpm. The interface between
the stationary and rotating domains was modeled using the multiple
Input power based on vortex height ðIHP2 Þ ¼ γhQ ð8Þ reference frame (MRF) method, which performs a local reference
frame transformation at the interface to enable flow variables in one
Efficiency based on total head ðηÞ ¼ ðBHP=IHPÞ × 100 ð9Þ zone to be used to calculate fluxes at the boundary of the adjacent
zone. Contrary to the arbitrary Lagrangian-Eulerian (ALE) method,
MRF is basically an approximation method in which the mesh does
Efficiency based on vortex height ðη2 Þ ¼ ðBHP=IHP2 Þ × 100 not deform or move. General grid interface (GGI) was used at the
ð10Þ interface of the stationary and rotating meshes. The GGI algorithm
performs an automatic surface trimming function, which can ac-
In case of any other turbine, the head H is the net head of the count for mismatch surfaces. The interface is constructed between
inlet; however, in case of GWVTs, the head is actually the height of the overlapping regions of the two sides of the interface (Ansys
the vortex because the water entry height above the vortex surface 2015). In view of the size of the reference basin, different velocities
does not result in any positive effect on the vortex. Therefore, η2 ranging between 0.1 and 0.5 m=s were selected at the inlet. For
was also calculated to measure the efficiency of the plant at any velocities greater than this range, the basin diameter should be in-
particular revolutions per minute. creased along with the outlet diameter. The water inlet velocity for
In basin analysis, only the boundary conditions of the basin do- the blade analysis was fixed at 0.45 m=s based on the results
main were used because no blade domain was observed. Therefore, obtained by testing the finalized basin at various velocities.
two fluids, namely, air and water, were defined for the analysis be- Multiphase domain was defined for all the analyses because
cause the air-core vortex resulted from the air–water interface. The water and air were present in the basin. The present numerical
properties of water and air at 25°C and 101.325 kPa (1 atm) pres- model involves a free surface between air and water (distinct
sure, respectively, were considered. Multiphase simulation was phases), which is required to be tracked across the computational

© ASCE 04018042-4 J. Energy Eng.

J. Energy Eng., 2018, 144(4): 04018042


domain by the virtue of the multifluid scheme. Therefore, the to the basin diameter lies in the range of 0.2–0.3. In a similar man-
volume of fluid (VOF) multiphase algorithm was employed ner, the effect of inlet channel height was investigated in terms of
in the present study to capture the sharp interface between two ratio of inlet height to basin height ranging from 0.1 to 0.5 (Table 1).
fluids that coexist in the computational domain without phase inter- An increase of this ratio causes the air core to diminish since the
penetration. The volume fraction uses α as a phase indicator to cap- excessive mass of water that enters the basin replaces the air core.
ture the interface between the two phases. A value of α ¼ 1 or 0 The measured parameters for this variation were also vortex height
shows that a control volume is entirely filled or empty, respectively, and gain in tangential velocity (Chattha et al. 2017). The effect of
with the primary phase, and a value between 1 and 0 indicates an outlet diameter was observed in the form of ratio of outlet diameter
interface. Precise modeling of the multiphase flow with the neces- to the basin diameter ranging from 0.13 to 0.17 (Table 1). Increas-
sary boundary conditions to capture the sharp air–water interface as ing this ratio allows more water to flow out of the basin; therefore, a
depicted in the water volume fraction contours is required. This can reduction in the water height in the basin was observed. In this case,
be achieved by using additional mathematical expressions to incor- therefore, the tangential velocity appeared to be approximately
porate the volume fraction in the momentum and mass conservation constant. The variation in maximum tangential velocity was also
equations to solve for the location of air and water in the computa- negligible because of the decrease in vortex height (Chattha
Downloaded from ascelibrary.org by Universitas Indonesia on 01/22/19. Copyright ASCE. For personal use only; all rights reserved.

tional domain et al. 2017). When the water height reduces, the friction offered
by the basin floor is more significant. The combined effects of these
X
N
modifications resulted in the formation of a full air core with a high
ρp ¼ ðαi ρi Þ ð11Þ
i¼1
vortex height. Moreover, the optimized basin resulted in 85.5% in-
crease in the tangential velocity than the initial or reference basin
X
N mainly because of the formation of the air core. Fig. 2(b) shows the
up ¼ ðαi ui Þ ð12Þ streamlines and the water–air interface in the optimized basin. The
i¼1 vortex that formed in this basin was fully developed and extended
from the top of the water surface down to the bottom outlet of the
where subscript p refers to phases (air and water). The provided
basin. Fig. 2(c) shows the separation between air and water in the
momentum and mass conservation Eqs. (1)–(4) were modified
form of volume fractions of air and water. The upper region rep-
by using the volume fraction equations for density and velocity
resents air, whereas the lower region represents the water present
across the computational domain. Atmospheric pressure and con-
inside the basin. The detailed results of the basin analysis have al-
ditions were applied at the top surface and the bottom drain of the
ready been reported in Chattha et al. (2017).
basin. The inlet was introduced with normal velocity of αair ¼ 0
and αwater ¼ 1, whereas the outlet was exposed to atmospheric
air with αair ¼ 1 and αwater ¼ 0, where αair is the volume fraction Blade Analysis
of air and αwater is the volume fraction of water. These input param-
eters were valid for both cases, namely, with and without turbine For the blade analysis, initialization of the rotating domain is re-
blades. These conditions can be physically understood by consid- quired as practiced by many researchers in the modeling and sim-
ering the turbine when it is not operational. Only water was present ulation of water turbines, such as Sammartano et al. (2013),
at the basin inlet. Consequently, air was only present at the outlet Khosrowpanah et al. (1988), and Andrade et al. (2011). In the first
and at the upper opening of the basin. The basin walls were set as step of the domain initialization, the domains were identified as
rough walls with sand-grain roughness of 0.045, which was the static or rotating. In the second stage the rotating domain was as-
sand-grain roughness of the steel sheet used for the experimental signed with the rotational speed of the runner based on the exper-
model. imental results as an input parameter. The same approach has been
followed in the present study to conduct the blade analysis. The
effects of blade placement in the selected basin were analyzed in
Results and Discussion of the Computational Study the second phase of the computational part of the study. Fig. 3(a)
shows the decrease in vortex height when the blades in the basin
rotate at 50 rpm. Fig. 3(b) shows the velocity streamlines formed in
Basin Analysis the presence of blades. The insertion of blades in the basin caused
The reference basin was analyzed in the initial phase of the com- distortion in the fluid flow. This phenomenon resulted in the de-
putational part of the study. The vortex did not form properly even crease in vortex height due to the decrease in the tangential velocity
at a mass flow rate of 5.5 kg=s because of the absence of an air of the fluid and a corresponding increase in the radial and axial
core, which is expected to generate small power [Fig. 2(a)]. Thus, velocities. The same can be observed in Fig. 3(c), which shows
in search of a suitable basin configuration, the basin geometric the volume fraction of water in the basin in the presence of blades.
parameters were varied until a suitable vortex was formed with The top surface in the basin and inlet channel represents the water–
the air core present throughout the vortex height. This basin was air interface and clearly shows the decrease in vortex height with
more suitable for a power generation that uses a mass flow rate the rotation of blades, thus showing a clear distinction from
of 4.5 kg=s. The basin was optimized by narrowing the inlet chan- Fig. 2(c), which had no presence of blades.
nel to smoothen the entry of water into the basin, increasing the The blades placed in the path of the flowing fluid produced
height of the inlet channel to allow the formation of a vortex with a rotational effect that was strongly influenced by the angular
high height, and increasing the size of the bottom outlet. The effect velocity. At small angular velocities, all blades experienced higher
of inlet channel width was investigated by using the dimensionless torques, which tend to decrease nonlinearly with the increase in
ratio of inlet channel width with the basin diameter ranging from angular velocities (Fig. 4). The nonlinearity of the graphs reflects
0.1 to 0.5 (Table 1). An increase in this ratio increases the mass flow the impact of variation in the vortex height caused by the blades at
into the basin, which causes the water height to rise until overflow various angular velocities. The cross-flow blades (i.e., Profile 2)
from the upper walls of the basin occurs. The analysis based on the exhibited the largest values of torque at any rotational speed among
variation of vortex height and gain in tangential velocity as reported all blade profiles. The shape of the cross-flow blades caused the
in Chattha et al. (2017) shows that the optimum ratio of inlet width high torque magnitudes, which resulted in lesser distortion in the

© ASCE 04018042-5 J. Energy Eng.

J. Energy Eng., 2018, 144(4): 04018042


Downloaded from ascelibrary.org by Universitas Indonesia on 01/22/19. Copyright ASCE. For personal use only; all rights reserved.

Fig. 2. Velocity streamlines and fluid flow (a) in the initial or reference geometry; and (b) in the selected basin. (c) Volume fractions of air and water in
the selected basin.

vortex profile and the extraction of the maximum kinetic energy of than any other blade, thereby resulting in lesser torques. The
the flowing water. The torque on the curved blades (i.e., Profile 3) twisted blades yielded higher torques than the conical blades be-
was the least in all rotational speeds because water was allowed to cause their contact area was greater than the latter.
strike the blades and move smoothly toward the center of the vortex The water flowed tangentially when no blades were installed
core. Consequently, the curved blades distorted the vortex more inside the vortex and resulted in a particular height of the vortex.

© ASCE 04018042-6 J. Energy Eng.

J. Energy Eng., 2018, 144(4): 04018042


Downloaded from ascelibrary.org by Universitas Indonesia on 01/22/19. Copyright ASCE. For personal use only; all rights reserved.

Fig. 3. Effect of the presence of rotating blades in the selected basin on (a) vortex formation; (b) velocity streamlines; and (c) water volume fraction.

The placement of the blade distorted the vortex profile and in- decreased and the tangential velocity of the water increased. Thus,
creased the radial and axial components of velocity while decreas- higher angular velocities resulted in a higher vortex height com-
ing the tangential components. This decrease in tangential velocity pared with lower angular velocities. Different blades distorted
caused a decrease in the height of the vortex. When the angular the vortex to a different value, thereby decreasing its height depend-
velocity of the blades increased, the distortion in the vortex profile ing on the flows of the fluid in the presence of these blades (Fig. 5).

© ASCE 04018042-7 J. Energy Eng.

J. Energy Eng., 2018, 144(4): 04018042


water toward the center of the air core after striking it, thus causing
a reduction in the vortex height. Although the twisted blades were
not along the shape of the vortex profile, these blades also kept the
water from flowing along the free vortex path, thereby resulting in
higher vortex height than the curved blades.
The best efficiency was obtained at almost midrange rotational
speeds between the maximum-load condition and no-load condi-
tion for all blades [Fig. 6(a)]. The blades placed in the path of
the circulating fluid in the basin produced a torque whose magni-
tude was strongly influenced by the rotational speed. The product
of the torque and the rotational speed contributes to the generated
Fig. 4. Effect of turbine rotational speed (rpm) on torques generated at power in an inverse combination as shown in Fig. 4. Therefore, the
the blades. efficiency of a turbine reflects a combined effect of the torque and
the rotational speeds, thus indicating the best efficiency points in
Downloaded from ascelibrary.org by Universitas Indonesia on 01/22/19. Copyright ASCE. For personal use only; all rights reserved.

the midrange rotational speeds for a given head and flow rate,
which has been the case for all the hydraulic turbines (Dixon
and Hall 2014). The values of efficiencies for cross-flow blades
were higher than other blades because the torques generated on
these blades are maximal. The higher efficiency of the cross-flow
blades may be attributed to the vortex-retaining capability of the
blades, which ultimately resulted in higher torques. The curved
blades were least efficient because their shape poorly extracted en-
ergy from the vortex and caused a considerable reduction in the
vortex height.
Chattha et al. (2017) showed that the height of a water inlet
above the vortex surface does not affect the vortex formation.
Therefore, the height of the vortex may be considered as the head
Fig. 5. Effect of turbine rotational speed (rpm) on vortex height. of the water and the efficiency can also be plotted in terms of the
height of the vortex rather than the total head of the inlet channel
[Fig. 6(b)]. When the efficiencies of the four blades, with respect to
the vortex height, were plotted against rotational speeds, the cross-
flow blades still yielded the best efficiencies. However, at lower
rotational speeds, the curved blades yielded better efficiencies than
their conical-type counter parts. A comparison of the two efficiency
graphs [Figs. 6(a and b)] found that no major difference was ob-
served in the trends of the efficiency curves. The difference lies in
the magnitudes of the efficiency and the location of the best effi-
ciency point. The magnitude difference is due to the difference in
the actual head and vortex height. For the case of conical and
curved blades, the best efficiency point is shifted toward the lower
rotational speeds, showing that they are more suitable for small
rotational speeds.

Validation of Numerical Results with Experiments

An experimental setup was fabricated and used to check the validity


of the numerical simulation results. The dimensions of the labora-
tory model were based on the available literature, and the setup was
then modified based on the findings of the numerical analyses to
form a strong air-core vortex for the mechanical power production.
In the case of the prototype, the circulation required for the gen-
eration of the vortex was achieved by the tangential entry of the
water into the cylindrical basin. The schematic of the experimental
Fig. 6. Effect of turbine rotational speed (rpm) on (a) turbine effi- setup is shown in Fig. 7. A pump provided water to the inlet chan-
ciency; and (b) turbine efficiency with regard to vortex height. nel and directed the water into the vortex pool. After the vortex
formation, the water exited through the bottom hole of the basin
and moved into the water reservoir, in which the water was recir-
culated by the pump. The flow rate was controlled by the ball valve
The cross-flow blades caused a minimum distortion to the vortex installed in the loop after the pump.
because of their shape, which guided the water to follow the vortex
profile. Similar vortex height was observed when the conical blades
Experimental Procedure
were analyzed because they were along the vortex profile. The de-
crease in vortex height caused by curved blades was greater than Flow rate was measured by calculating the dynamic pressure using
any other blades because of their curved shape, which directed the two pressure gauges. These gauges individually measured the static

© ASCE 04018042-8 J. Energy Eng.

J. Energy Eng., 2018, 144(4): 04018042


Downloaded from ascelibrary.org by Universitas Indonesia on 01/22/19. Copyright ASCE. For personal use only; all rights reserved.

Fig. 7. GWVT experimental setup.

and stagnation pressures. Dynamic pressure was obtained by the


difference between the two gauge readings. This dynamic pressure
was used to calculate the water velocity, which was further multi-
plied by the area of the pipe to obtain the volumetric flow rate. The
flow rate was maintained at 4.2 L=s. A domestic bucket of 30-cm Fig. 8. Experimental results for all blades: (a) torque and efficiency;
height and 35-cm diameter was graduated at 22 L to validate the and (b) torque and efficiency with regard to vortex height.
calculated volumetric flow rate through the pump. The time taken
by the pump to fill the bucket was measured using a stopwatch. The
basin was graduated at different levels to measure the vortex height.
Table 3. Comparison between numerical and experimental efficiencies
A coupler was attached to the top of the output shaft, which was
connected to the blades. The radius of the coupler was 19 mm. Difference between numerical
Spring balances were used to apply force to the output shaft. The and experimental results
torque applied on the shaft was obtained when this force, which Blade Efficiency 1 (%) Efficiency 2 (%)
was measured by spring balances, was multiplied by the radius of Profile 1 7.06 7.92
the coupler. A digital tachometer was used to measure the angular Profile 2 9.31 12.28
velocity of the output shaft. Profile 3 5.05 6.05
Profile 4 9.17 10.46

Experimental Results
The increase in the resisting torque on the coupler caused an in-
crease in the load on the turbine. When the load on the turbine in- of a GWVT, then all blades have comparable efficiencies for small
creased, the angular velocity of the blades decreased, causing a torques.
change in the efficiency of the plant. The efficiency increased until
the vortex distorted significantly due to flow disturbance. Fig. 8(a)
Comparison of Experimental and Simulation Results
shows that the optimum torque, in which the highest efficiency is
obtained, lies halfway between no force on the blade and the maxi- The angular velocity of the blades decreased when the force applied
mum force to stop the blade. The performance trend of the blade on the blades was increased. The sudden decrease in the angular
was maintained in the case of numerical simulation. Cross-flow velocity was due to the decrease in vortex height, which implies
blades outclassed other profiles in terms of efficiency. By contrast, the decrease in vortex strength. This decrease became drastic when
the curved blades were the least performer. The results recommend the torques significantly increased. The results obtained from
the use of cross-flow-type blades in a GWVT, with an experimental ANSYS CFX simulations showed close resemblance to the results
efficiency approaching 50%. obtained experimentally. Furthermore, an error that ranges from 6
The difference among the performances of the four blade pro- to 15% was observed in different cases. Fig. 9(a) indicates the dif-
files was small when the efficiency of blades was plotted using the ference observed between numerical and experimental results in
vortex height as the total head. In this case, the cross-flow blades one of the blades. The power obtained numerically was slightly
were more efficient than other blades with increased efficiencies in higher than the power obtained experimentally because mechanical
all cases. Although the efficiency difference was negligible for all frictional forces were not considered in the numerical analysis.
blade profiles with small values of the applied torques, the differ- Moreover, at higher rotational speeds, the centrifugal forces caused
ence became significant when higher resisting torques were applied an increase in the height of the vortex that resulted in higher torques
on the blades [Fig. 8(b)]. The highest efficiency increase was ob- on the blades. Furthermore, the use of rotational speed as an input
served for the case of conical blades and the least efficiency was parameter resulted in higher numerical results. However, the blade
observed in the case of cross-flow blades. Table 3 shows the differ- performance trend shown in the experimental results was similar to
ence between the numerical and the experimental results in the ef- the results in the numerical simulation. Cross-flow blades had the
ficiency for all the blades at their best efficiency points. The results highest power output among any other types of blades, which
highlight the importance of vortex height in measuring the perfor- shows their vortex-retaining capability. By contrast, the curved
mance of GWVT. If the vortex height was maintained in the basin blades had the least power output among all cases because they

© ASCE 04018042-9 J. Energy Eng.

J. Energy Eng., 2018, 144(4): 04018042


Downloaded from ascelibrary.org by Universitas Indonesia on 01/22/19. Copyright ASCE. For personal use only; all rights reserved.

Fig. 9. Comparison of numerical and experimental results in all blades based on (a) output power; and (b) vortex heights.

distorted the vortex. The results recommend the use of cross- causing great frictional loss in the water velocity. Water entry
flow-type blades in a GWVT with an experimental efficiency ap- above the upper surface of the vortex does not result in any pos-
proaching 50%. itive output.
Fig. 9(b) shows that the increase in torques applied on the blades The finalized basin configuration has been used in the blade
decreases angular velocity, which correspondingly decreases the analysis with different blade shapes at various load conditions.
vortex height. The decrease in vortex height caused by cross-flow The blade analysis suggests that cross-flow blades are the most suit-
blades was minimal because they maintain the water motion along able blades for GWVT. Vortex height has been proven to be an
the vortex flow direction. The decrease in vortex height was maxi- important measure of the performance of GWVT. One of the lim-
mal in the case of the curved blades because they changed the itations of this study is that the weight of the blades is not consid-
direction of the water striking the blade edges from tangential to ered in the numerical simulation, which increases the difference
radial. The difference between the numerical and experimental between the experimental and numerical results with the increase
results increased at higher angular velocities because of the meth- in blade weight. The height of the vortex decreases when the load
odology used for the numerical analysis of the blade. In the simu- on the turbine is increased. Based on the study, the proposed meth-
lations, the blade rotation was an input parameter. At higher angular odology may also be used to determine the plant specifications and
velocities, the centrifugal forces came into play and the blades blade size and shape for various flow rates and heads.
tended to cause an outward flow of the water, thereby reducing
the radial velocity and increasing the vortex height.
Appendix. Mesh Sensitivity Analysis

Conclusions Vortex height was calculated using three different inlet velocities
to conduct mesh sensitivity analysis of the basin as shown in
GWVT is an efficient turbine that can be used as a microhydro Fig. 10(a). Similarly, mesh sensitivity analysis for four different
turbine to generate power. Based on the findings of the study, a blade profiles was conducted by computing torque at fixed rota-
two-phase CFD-based study is always important for the analysis tional speed of 110 rpm as shown in Fig. 10(b). For further analysis
of a GWVT. The configuration with a fully developed air core of the present study, the combination of mesh elements was se-
throughout the vortex height is the most suitable for the basin. lected based on the fact that both vortex height and torque tend
When a basin with a large diameter is used, the vortex height de- to remain unchanged despite further increase in the number of mesh
creases and almost all the water glides along the floor of the basin, elements.

© ASCE 04018042-10 J. Energy Eng.

J. Energy Eng., 2018, 144(4): 04018042


References
Andrade, J. D., C. Curiel, F. Kenyery, O. Aguill, A. Vasquez, and
M. Asuaje. 2011. “Numerical investigation of the internal flow in
a Banki turbine.” Int. J. Rotating Mach. 2011: 1–12. https://doi.org/10
.1155/2011/841214.
Ansys. 2015. ANSYS CFX-solver modeling guide Release 15.0.
Canonsburg, PA: Ansys.
Chattha, J. A., T. A. Cheema, and N. H. Khan. 2017. “Numerical inves-
tigation of basin geometries for vortex generation in a gravitational
water vortex power plant.” In Proc., 8th Int. Renewable Energy
Congress (IREC). Amman, Jordan: IEEE.
Chen, Y., C. Wu, B. Wang, and M. Du. 2012. “Three-dimensional numeri-
cal simulation of vertical vortex at hydraulic intake.” Proc. Eng. 28:
55–60. https://doi.org/10.1016/j.proeng.2012.01.682.
Downloaded from ascelibrary.org by Universitas Indonesia on 01/22/19. Copyright ASCE. For personal use only; all rights reserved.

Dhakal, S., S. Nakarmi, P. Pun, A. B. Thapa, and T. R. Bajracharya. 2014.


“Development and testing of runner and conical basin for gravitational
water vortex pool.” J. Inst. Eng. 10 (1): 140–148. https://doi.org/10
.3126/jie.v10i1.10895.
Dhakal, S., A. B. Timilsina, R. Dhakal, D. Fuyal, T. R. Bajracharya, H. P.
Pandit, N. Amatya, and A. M. Nakarmi. 2015. “Comparison of cylin-
drical and conical basins with optimum position of runner: Gravitational
water vortex power plant.” Renewable Sustainable Energy Rev. 48 (8):
662–669. https://doi.org/10.1016/j.rser.2015.04.030.
Dixon, S. L., and C. A. Hall. 2014. Fluid mechanics and thermodynamics
of turbomachinery. 7th ed. 367–369. Amsterdam, Netherlands: Elsevier.
Elbatran, H., O. B. Yaakob, Y. M. Ahmed, and H. M. Shabara. 2015.
“Operation, performance and economic analysis of low head micro-
hydropower turbines for rural and remote areas: A review.” Renewable
Sustainable Energy Rev. 43 (3): 40–50. https://doi.org/10.1016/j.rser
.2014.11.045.
Fig. 10. Mesh sensitivity analysis for (a) basin; and (b) blades. Khan, M. J., M. T. Iqbal, and J E. Quaicoe. 2008. “River current energy
conversion systems: Progress, prospects and challenges.” Renewable
Sustainable Energy Rev. 12 (8): 2177–2193. https://doi.org/10.1016/j
.rser.2007.04.016.
Acknowledgments Khosrowpanah, S., A. A. Fiuzat, and M. L. Albertson. 1988. “Experimental
study of cross flow turbine.” J. Hydraul. Eng. 114 (3): 299–314. https://
This study was supported by a National Research Foundation of doi.org/10.1061/(ASCE)0733-9429(1988)114:3(299).
Korea (NRF) grant funded by the Korea government (MSIP) Kueh, T. C., S. L. Beh, D. Rilling, and Y. Ooi. 2014. “Numerical analysis of
(No. 2017R1A2B2005515) and a grant from the Priority Research water vortex formation for the water vortex power plant.” Int. J. Inno-
Centers Program through the NRF, as funded by the MEST vation Manage. Technol. 5 (2): 111–115. https://doi.org/10.7763/IJIMT
(No. 2010-0020089). The authors also acknowledge the support .2014.V5.496.
from the Pakistan Science Foundation (PSF) for this study under Li, H., H. Chen, M. A. Zheng, and Z. Yi. 2008. “Experimental and numeri-
Research Project No. PSF/ILP/KPK-GIKI/Engg(073). cal investigation of free surface vortex.” J. Hydrodyn. Ser. B. 20 (4):
485–491. https://doi.org/10.1016/S1001-6058(08)60084-0.
Mohanan, A. 2016. “Power generation with simultaneous aeration using
Notation a gravity vortex turbine.” Int. J. Sci. Eng. Res. 7 (2): 19–23.
Odgaard, J. 1986. “Free-surface air core vortex.” J. Hydraul. Eng. 112 (7):
The following symbols are used in this paper: 610–620. https://doi.org/10.1061/(ASCE)0733-9429(1986)112:7(610).
g = gravitational acceleration (m2 =s); Power, C., A. McNabola, and P. Coughlan. 2015. “A parametric experimen-
tal investigation of the operating conditions of gravitational vortex hy-
H = total head (m);
dropower.” J. Clean Energy Tech. 4 (2): 112–119. https://doi.org/10
h = vortex height (m); .7763/JOCET.2016.V4.263.
N = rotational speed (rpm); Sammartano, V., C. Aricò, A. Carravetta, O. Fecarotta, and T. Tucciarelli.
Q = flow rate (m3 =s); 2013. “Banki-Michell optimal design by computational fluid dynamics
r = radius (m); testing and hydrodynamic analysis.” Energies 6 (5): 2362–2385. https://
T = torque (N-m); doi.org/10.3390/en6052362.
Shabara, H. M., O. B. Yaakob, Y. M. Ahmed, A. H. Elbatran, and M. S. M.
V r = radial velocity (m=s);
Faddir. 2015. “CFD validation for efficient gravitational vortex pool
V z = axial velocity (m=s); system.” J. Teknol. (Sci. Eng.) 74 (5): 97–100. https://doi.org/10
V θ = tangential velocity (m=s); .11113/jt.v74.4648.
αair = air volume fraction; Singh, P., and F. Nestmann. 2009. “Experimental optimization of a free
αwater = water volume fraction; vortex propeller runner for micro hydro application.” Exp. Ther. Fluid
γ = specific weight of the water (N=m3 ); Sci. 33 (6): 991–1002. https://doi.org/10.1016/j.expthermflusci.2009
.04.007.
η = efficiency;
Singh, P., and F. Nestmann. 2011. “Experimental investigation of the in-
ν = kinematic viscosity (m2 =s); fluence of blade height and blade number on the performance of low
ρ = density (kg=m3 ); and head axial flow turbines.” Renewable Energy 36 (1): 272–281. https://
ω = rotational speed of the rotor (rad=s). doi.org/10.1016/j.renene.2010.06.033.

© ASCE 04018042-11 J. Energy Eng.

J. Energy Eng., 2018, 144(4): 04018042


Venukumar, A. 2013. “Artificial vortex (ArVo) power generation: An Wanchat, S., R. Suntivarakorn, S. Wanchat, K. Tonmit, and P. Kayanyiem.
innovative micro hydroelectric power generation scheme.” In Proc., 2013. “A parametric study of gravitational water vortex power plant.”
Global Humanitarian Technology Conf.: South Asia Satellite Adv. Mater. Res. 805 (9): 811–817. https://doi.org/10.4028/www
(GHTC-SAS), 53–57. New York: IEEE. .scientific.net/AMR.805-806.811.
Wanchat, S., and R. Suntivarakorn. 2012. “Preliminary design of vortex Wang, Y., C. Jiang, and D. Liang. 2010. “Investigation of air-core vortex at
pool for electrical generation.” Adv. Sci. Lett. 13 (1): 173–177. https:// hydraulic intakes.” J. Hydrodyn. Ser. B. 22 (5): 696–701. https://doi.org
doi.org/10.1166/asl.2012.3855. /10.1016/S1001-6058(10)60017-0.
Downloaded from ascelibrary.org by Universitas Indonesia on 01/22/19. Copyright ASCE. For personal use only; all rights reserved.

© ASCE 04018042-12 J. Energy Eng.

J. Energy Eng., 2018, 144(4): 04018042

You might also like