Anagnos To Poulos 2012

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

See

discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/267494263

Optimal Design and Experimental Validation of a


Turgo Model Hydro Turbine

Conference Paper · July 2012


DOI: 10.1115/ESDA2012-82565

CITATIONS READS

9 534

4 authors, including:

Phoevos K. Koukouvinis Dimitris E. Papantonis


City, University of London National Technical University of Athens
31 PUBLICATIONS 119 CITATIONS 66 PUBLICATIONS 860 CITATIONS

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

FS3000 View project

Nozzle flow simulations with multi-component fuels and additives View project

All content following this page was uploaded by Phoevos K. Koukouvinis on 17 February 2016.

The user has requested enhancement of the downloaded file.


Proceedings of the ASME 2012 11th Biennial Conference on Engineering Systems Design and Analysis
ESDA2012
July 2-4, 2012, Nantes, France

ESDA2012-82565

OPTIMAL DESIGN AND EXPERIMENTAL VALIDATION OF A TURGO MODEL HYDRO TURBINE

John S. Anagnostopoulos
School of Mechanical Engineering, National Technical University of Athens, Greece

Phoevos K. Koukouvinis Fotis G. Stamatelos Dimitris E. Papantonis


School of Mechanical Engineering, National Technical University of Athens, Greece

ABSTRACT head multi-jet Pelton or the high head Francis turbines [1]. Like
This work presents the development and application of a new Pelton, Turgo turbine has a flat efficiency curve and provides
optimal design methodology for Turgo impulse hydro turbines. excellent part-load efficiencies hence it constitutes the best
The numerical modelling of the complex, unsteady, free surface solution for large flow rate variations. An additional advantage
flow evolved during the jet-runner interaction is carried out by a of Turgo turbine is that it can operate for long periods and
new Lagrangian particle method, which tracks a number of minimum wear when the water is laden with slit and other
representative flow elements and accounts for the various entrained matter. Also, larger jet and flow rates can be treated
hydraulic losses and pressure effects through special adjustable compared to a Pelton runner of same diameter. As a result, a
terms introduced in the particle motion equations. In this way, Turgo turbine has higher specific speed and smaller size than a
the simulation of a full periodic interval of the flow field in the Pelton turbine for the same power. Finally, since the water jet
runner is completed in negligible computer time compared to the enters one side of the runner and exits through the other, the
corresponding needs of modern CFD software. Consequently, interference of the outflow with the incoming jet can be minimal.
the numerical design optimization of runner geometry becomes In spite of the above advantages, Turgo turbines are much
feasible even in a personal computer and affordable by small and less spread than Pelton, mainly because the runner is more
local manufacturers. The bucket shape of a 70 kW Turgo model difficult to fabricate by local manufacturers: The buckets are
is properly parameterized and numerically optimized using a complex in shape and more fragile than Pelton buckets [2].
stochastic optimization software to maximize the hydraulic Moreover, the hydraulic efficiency is more sensitive on the exact
efficiency of the runner. The optimal runner and the rest turbine shape of the buckets surface than in Pelton. Therefore,
parts are then manufactured and installed in the Lab for testing. implementation of computational fluid dynamics in the design of
Detailed performance measurements are conducted and the modern Turgo turbines appears to be necessary in order to
results show satisfactory agreement with the numerical improve their efficiency and cost-effective construction beyond
predictions, thus validating the reliability and effectiveness of the traditional design practices.
the new methodology. Very few scientific articles can be found in the literature
Keywords: Turgo impulse turbine; Particle simulation method; dealing with the design of Turgo runner, and only few
Bucket parameterization; Design optimization; Performance companies manufacture this turbine type worldwide. An
measurements extended description is first given in Gibson [3], but without any
information about dimensioning and design of the runner. Some
details on the latter can be found in [4].
1. INTRODUCTION All published works on impulse hydroturbines concern the
most common Pelton type turbine. However, the distribution
The Turgo impulse hydroturbine can be used in medium to system, the spear valve injector, and the impacting jet flow are
high heads, from 15 to 300 m, and it was first patented in 1920. common to both types. The importance of jet quality and its
Its design and efficiency are remarkably improved during the relationship with turbine efficiency has been thoroughly
next decades, rendering it capable of competing with the low- investigated in [5-7], and the results can be applied also to Turgo

1 Copyright © 2012 by ASME

Downloaded From: http://proceedings.asmedigitalcollection.asme.org/ on 02/01/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use


turbines. On the other hand, detailed measurements of the free- angular location of the bucket (θo, Fig. 1). Calculations are then
surface flow properties in the buckets are very difficult because continued for the particles of oncoming frames, until all particles
of the contaminating effects of the water outflow. of a frame are blocked by the next coming buckets (jet cut).
Thanks to the continuous increase of computing power and to
the development of advanced simulation methods, complete
analysis of the complex unsteady, free-surface flow developed
during the jet-runner interaction in impulse turbines is possible,
though several complex secondary flow mechanisms are still not
modelled, and some inaccuracies associated with the turbulence
modeling or the boundary conditions, cannot be avoided. Several
such studies have been published, but only for Pelton runners.
The volume-of-fluid method and the two-phase
homogeneous or inhomogeneous models are implemented in
various Eulerian solvers [8-11]. The traditional mesh-based
Eulerian approaches face significant numerical diffusion
problems due to the complex evolution of the free surface flow
pattern. Mesh-less particle simulation approaches can be
advantageous in such flows, due to the inherent free surface
modelling. The Smoothed Particle Hydrodynamic appears to be
the most promising such method, and a recent hybrid SPH-ALE
approach [12] can help to overcome the main drawbacks of Fig. 1. Initial distribution of representative flow particles.
standard SPH related to stability and accuracy. The SPH method
has also applied by the present authors in a first attempt for a
CFD simulation of flow field in a Turgo runner [13, 14]. The equations of motion of the fluid particles are solved in a
These computations can simulate the entire working cycle of rotating orthogonal system of reference, and are expressed as
the bucket, but the accuracy of the torque predictions is still not follows:
adequate, mainly due to the development of complex secondary
flow mechanisms during the jet impingement and cut. Also, due d 2x
to the unsteady nature of the flow field, the above solvers require  f x x, y 
dt 2
large computational effort for a single solution. Hence, a
d2y dz
complete design optimization of an impulse turbine runner,  f y  x, y    2 y  2  (1)
which may need thousands of flow evaluations, requires huge dt 2 dt
computing power and remains non-feasible for industrial design. 2
d z dy
This paper presents the formulation and application results of 2
 f z x, y    2 z  2 
dt dt
a new Fast Lagrangian Simulation (FLS) method, which is
developed for the numerical design optimization of impulse
turbine runners at minimal computer cost. The method has been where x, y and z are the Cartesian coordinates as defined in Fig.
recently used to study both Pelton and Turgo runners design [15, 1, ω is the angular rotation speed of the runner and fx, fy, and fz
16], and in the present work it is applied in order to design and are functions of the local surface geometrical characteristics.
manufacture a prototype Turgo model, which is then tested in the
Lab.

2. FLS MODEL FORMULATION

2.1. Flow modeling

The numerical simulation of the fluid flow on the bucket


surface is based on the Lagrangian approach, and the trajectories
of an adequate number of representative fluid particles are
tracked in order to produce statistically accurate results. The jet
volume is divided into several consecutive segments or frames,
and a number of particles are uniformly distributed over the
circular area of each frame, as shown in Fig. 1.
The jet-bucket interaction starts when at least one particle of Fig. 2. Indicative fluid particle trajectories
a frame impinges on the inner bucket surface, at a certain

2 Copyright © 2012 by ASME

Downloaded From: http://proceedings.asmedigitalcollection.asme.org/ on 02/01/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Calculation of a particle trajectory is divided into two parts
(Fig. 2). At first, the standard particle equations without the f- Flow spreading: In order to model the pressure effects on
functions above are solved to reproduce the particles path in the the surface flow evolution, each particle acquire at the impact
free jet, and the algorithm checks only if the particle impinges on point an artificial “spreading” velocity component perpendicular
the next coming bucket in order to stop the integration.

to its impacting plane ( VS , Fig. 3), while its main velocity
Otherwise, the tracking process continues until the particle
encounters the inner bucket surface or passes by the bucket component is correspondingly reduced to preserve kinetic
without impinging. A third option is also possible in some off- energy. The finally adopted scheme involves two more
design operating conditions, when a particle may hit on the hub adjustable coefficients to compute the magnitude of the
or the shroud of the runner. “spreading” velocity. The first quantifies the influence of the
The particle tracking after impacting on the bucket is relative position of a particle in the jet (radial and angular
performed by numerical integration of Eqs. (1), assuming that location in respect to the jet axis). The second is introduced to
the particle slides along the inner 3-dimensional surface (Fig. 2) account for a possible effect of the initial jet diameter (or the
of known geometric characteristics (local depth and slopes). A spear valve opening) on its spreading rate after the impact.
second-order predictor-corrector scheme is adopted for the step-
by-step calculation of the trajectory and the procedure is
repeated until the particle flows out from the bucket.
The above equations do not contain particle interaction or
mechanical losses terms and hence they cannot reproduce the
real flow picture in the bucket. For this reason, the FLS model
introduces a number of additional terms in order to account for
the hydraulic losses along the particle’s pathway, and for the free
surface flow spreading rate in the bucket.

Friction losses: Assuming a constant mean friction


coefficient, the kinetic energy of a particle reduces by a factor
analogous to the square of particle velocity and to the sliding
distance, hence the new particle velocity magnitude after a time
step Δt becomes:


V p'  V p  1  C f  V p   t  (2)
Fig. 3. Modeling of the flow spreading on the bucket surface
where Cf is a friction-loss adjustable coefficient.

Impact losses: Additional kinetic energy losses take place The performance and reliability of the FLS model depend on
due to abrupt change of particle momentum at its impact point. the appropriate tuning of the values of the above coefficients and
The kinetic energy losses are taken analogous to the square of this task can be accomplished with the aid of experimental data
the normal, to the surface, particle velocity component, and this or based on numerical results obtained by a more accurate CFD
gives: solver.

V p '  V p  1  C i  cos 2  i  (3) A sufficiently small time step (2·10–5 sec) is used for the
numerical integration of the particle motion equations in order to
eliminate numerical error. A total number of the order of 103
where Ci is the impact-loss adjustable coefficient and φi the trajectories was found adequate to produce independent results
angle between the particle impingement velocity and the unit for the hydraulic efficiency of the runner, whereas for the
vector normal to the surface at the impact point. reproduction of the surface flow pattern more than 104
trajectories were computed. Even in the latter case, the CPU time
Change direction: The progressive change of particle’s path requirements are very small, just a few seconds in a modern
direction (and momentum) as it slides along the curved bucket personal computer.
surface also causes minor energy losses, which can be modeled
using a similar to the impact losses term:
2.2. Runner geometry and parameterization

V p '  V p  1  C p  cos 2   (4)
The mean velocity of the free jet emerging from the nozzle of
the turbine is determined from the net head, by the equation:
where Cp is the adjustable coefficient and Δφ the angular change
in direction of the sliding particle during the time step Δt.

3 Copyright © 2012 by ASME

Downloaded From: http://proceedings.asmedigitalcollection.asme.org/ on 02/01/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use


V jet    2 g H  0,97  2 g H (5)
The average bucket inlet and outlet angles, β1 and β2, can be
computed from the corresponding velocity triangles (Fig. 4b). At
where the velocity coefficient  stands for the hydraulic losses in the best efficiency point the flow exits with almost zero
the spear valve, and it is usually between 0.95 and 0.99 [17]. The circumferential velocity, hence the outlet velocity triangle is
value of  =0.97 is taken for the present study. The orthogonal and can also be constructed (Fig. 4b). However, due
corresponding jet diameter, Djet, can be obtained from the to rotation, the jet reaches a bucket at different locations:
nominal flow rate: Initially closer to the hub and then progressively towards the
shroud. Consequently, the velocity triangles defined in Fig. 4b

QK   D 2jet  V jet (6) are not representative of the whole jet-bucket interaction. For
4 this reason, additional design parameters are introduced to allow
for differentiation of the inlet and outlet bucket angles along the
At the best efficiency point in impulse turbines the leading and trailing edges of the bucket. Also, the curvature of
circumferential speed, u, of the runner is about half of the jet these lines can be varied. The above make a total of 12
velocity. Hence the pitch diameter of the runner is (Fig. 4a): geometric control variables for parametric description of the
bucket geometry.
60 u The mean 3-dimensional surface of the bucket is then
Ds  (7)
n generated using the conformal mapping methodology and
interpolation techniques. An example of the resulting shape is
where n is the rotation speed in rpm. shown in Fig. 5. The above parameterization method ensures
The runner has a conical shape in the meridian plane (Fig. always a smooth curvature variation of the surface. The inner
4a). The inlet bucket edge is a straight line and the inlet width is and outer surfaces of the bucket can be constructed considering a
larger than the jet diameter (b1 ≈ 1,2Djet [4]), in order to secure given bucket thickness distribution. Finally, the hub and the
the entrance of the entire jet even for the highest flow rate. The shroud are easily introduced as axisymmetric surfaces, and the
outlet edge of the bucket is drawn here with the aid of a Bezier entire runner can be reproduced as shown in Fig. 6.
curve. The bucket traces on the hub and the shroud, as well as an
intermediate streamline are also generated using corresponding
Bezier polynomials. Hence a number of Bezier control points are
introduced as design parameters.

Ds
(a)
b1

B Fig. 5. Indicative view of mean bucket surface.

70-80 o

c1=c w1
β1
β1
u1

(b) β2 w2
c2 β2

u2
.
Fig. 4. Turgo runner configuration:
a) Meridian plane; (b) Velocity triangles. Fig. 6. Turgo runner drawn by the present method.

4 Copyright © 2012 by ASME

Downloaded From: http://proceedings.asmedigitalcollection.asme.org/ on 02/01/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use


The mechanical torque developed on the runner is computed
2.3. Monitoring and post processing the results from the following equation of conservation of angular
momentum:
The FLS model includes post-processing of the results in
order to compute the turbine performance variables, such as the
developed torque on the buckets and the hydraulic efficiency of

M run   Qu rin win  rout wout  (8)

the runner. Also, it can calculate the local forces exerted on the
where Qu is the cumulative flow rate that enters each bucket and
bucket due to the change of fluid particles momentum, and hence
w the absolute tangential velocity component at corresponding
to assess the local contributions to the energy conversion. Snap-
radial distance r. Subscripts in and out denote the time instants
shots of the flow pattern at any time instant can be easily
when a particle impinges on or leaves the bucket, respectively
obtained, and can be used to produce animation in graphics
(Fig. 2). The mean angular momentum at the inlet, assuming a
software at very low computer cost.
uniform jet velocity Vjet, becomes:
Indicative such flow pictures are illustrated in Fig. 7. The
free-surface flow on a bucket starts, evolves and terminates
within about 110-120 degrees of runner rotation. Soon after its rin win  Rrun V jet  cos  (9)
impact on the reference bucket (Fig. 7a), the jet starts to interact
with the next coming bucket too (not shown in Fig. 7), which where Rrun is the runner pitch radius and θ the jet angle relative
eventually cuts the jet (Fig. 7c). A noteworthy behavior that can to the runner disk (Fig. 4b). The mean angular momentum at the
be observed is that the flow leaves the bucket from quite bucket outlet is computed by averaging the local fluid particle
different regions during the interaction period (Figs. 7b,c,d), properties monitored there:
because of the bucket displacement due to rotation. Therefore,
the correct design of the entire trailing edge line is decisive in 1
order to minimize the kinetic energy of the outflow and thus to rout wout 
N
 ri wi (10)
achieve high hydraulic efficiency. i

where ri and wi are the radial distance and the absolute tangential
(b) velocity component of a particle i at the moment it flows out
(a) from the bucket, and N is the total number of fluid particles that
interact with a single bucket. The hydraulic efficiency of the
runner can then be obtained as the ratio of the developed
mechanical power divided by the corresponding net hydraulic
power at the inlet:

 run   run 
h  
 g Q H jet  g Q  2 H   (11)

where Hjet and H are the hydraulic head of the discharged jet and
the net head of the flow at the injector inlet, respectively, and Q
(d) is the nozzle flow rate. It must be noted that the runner flow rate
Qu in Eq. (8) may be less than Q in Eq. (11) for certain off-
design operating conditions, when some particles do not impinge
in the buckets.
(c)

2.4. Adjustment of model coefficients

The values of the FLS model coefficients included in Eqs. (1)


have been previously regulated with the aid of experimental data
taken from an 80 kW laboratory Pelton turbine model [15]. In
order to check their validity for the Turgo runner geometry, the
Fig. 7. Selected flow pictures: a) start of jet-bucket interaction; b) FLS solution is compared with the results of a more accurate but
full jet impingement; c) jet cut and end of impingement; d) much more expensive CFD simulation software, which is based
evacuation phase.
on the Smoothed Particle Hydrodynamic method [13, 14] (Fig.
8).

5 Copyright © 2012 by ASME

Downloaded From: http://proceedings.asmedigitalcollection.asme.org/ on 02/01/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use


The comparison showed certain differences, mainly with the aid of literature data [3, 4] and by present calculations,
concerning the spreading rate of the surface flow in the bucket. and are given in Fig. 10. The main design and operation data are
Hence, after several comparative flow evaluations the FLS tabulated in Table 1.
model coefficients are properly re-adjusted to achieve the best
possible agreement. The final results are substantially improved,
as shown in the indicative views of Fig. 9, at two bucket angular
positions after start of jet-bucket interaction (Δθ, in Fig. 1)

Fig. 10. Dimensions of the model runner.

Table 1. Main Turgo model characteristics.


Net head 48 mWG
Number of injectors 2
Fig. 8. Indicative picture of the jet-runner interaction simulated by
Nominal flow (per nozzle) 0,09 m3/sec
the SPH model.
Nominal Power (per nozzle) ~35,4 kW
Rotation speed 1000 rpm
Jet diameter (nominal) 62 mm
Jet angle 25 deg
Pitch diameter 285 mm

+36o
3.2. Runner geometry and design optimization
SPH FLS
A general stochastic optimization software, developed and
brought to market by the Lab of Thermal Turbomachinery
NTUA [18], is used in the present work to find the runner design
that achieves maximum hydraulic efficiency. The optimizer is
+72o based on evolutionary algorithms and it is suitable for complex
non-linear and multi-parametric problems, as the shape
optimization of complex 3D surfaces.
The algorithm selects values of the free design parameters
within the prescribed ranges and looks automatically for the set
Fig. 9. Comparison of free surface flow evolution results of the FLS that maximizes the cost function (here the hydraulic efficiency of
and the SPH models. the runner) using populations of candidate solutions. The
passage from a population to the next one that contains improved
solutions mimics the biological evolution of species generations
[19].
3. MODEL APPLICATION AND VALIDATION The fast flow evaluation achieved by the FLS model allows
for using wide variation limits for all design variables. As a
3.1. Turgo model specifications result, much different shapes are created and numerically solved
during the optimization procedure, as the examples of Fig. 11.
In order to validate the developed simulation methodology Due to the stochastic nature of the operation there is no strict
and assess its capability to be used for Turgo turbines design, a convergence criterion; the procedure is terminated when the
70 kW model turbine is designed, manufactured and tested in the efficiency does not increase further within a predetermined
Laboratory. The main dimensions of the runner are obtained number of consecutive evaluations.

6 Copyright © 2012 by ASME

Downloaded From: http://proceedings.asmedigitalcollection.asme.org/ on 02/01/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use


and nozzles are increased by about 30%, in order to be able to
measure the effect of larger jet diameters on the same runner. A
plexiglass window is opened at the front cover of the casing to
permit visual observation of the runner outflow (Fig. 15). The
new model is installed in one of the test rigs of the Laboratory of
Hydraulic Turbomachines, NTUA, after proper adjustments. A
simplified sketch of the rig is given in Fig. 16.
The characteristics of the main devices used for testing are
the following:
— Three-stage centrifugal pump, driven by a 220 kW AC
motor. The pump’s rotational speed can be adjusted through
the hydraulic transmission. At the BEP the pump rotates at
1800 rpm, providing 320 m3/h flow rate and 117 mWG net
head.
— 110kW DC generator, controlled by either rotational speed
or torque.
The turbine net head is adjusted by regulating the pump
speed, while the water flow rate is manually regulated from the
Fig. 11. Various bucket shapes tested during optimization.
spear valves by precision screw. The turbine rotational speed can
be adjusted through the DC generator, so it is possible to test the
A typical convergence curve is given in Fig. 12. The multi- turbine for a wide range of operating conditions.
parametric optimization requires several hundred evaluations
(flow field simulations) for convergence, but with the FLS solver
the whole procedure can be completed in only a couple of hours
in a modern PC.

3.3. Model construction and installation

The optimally designed runner of the Turgo model is


constructed by separate fabrication of its parts. The hub and
shroud are made by precision machining, and proper slots are
opened on the hub surface to accommodate the buckets (Fig. Fig. 13. Turgo model runner: a) buckets assembly; b) finished.
13a). A prototype bucket of composite material is constructed at
first by 3D-printing, and it is then used to produce aluminium
prototype. The runner buckets are fabricated by bronze casting
and finally assembled with bronze welding (Fig. 13b). This
procedure enhances the axial symmetry of the runner and
ensures that all buckets are identical.
0.9
Overall efficiency, η (%)

0.88

0.86

0.84 Fig. 14. Indicative spear valve drawing.

0.82
The performance data of the turbine model are obtained in
0.8 the form of characteristic operation curves of net head, shaft
1 10 100 1000
Number of evaluations power and overall efficiency, as function of the flow rate and
rotation speed. The corresponding physical quantities were
Fig. 12. Convergence history of the design optimization procedure. measured with the following instrumentation:
1) The Rotational Speed of the turbine by an electronic pulse
Detailed engineering drawings of all turbine model parts are meter with measurement error ±0.25 %.
created in CAD environment (Fig. 14). The spear valve and 2) The Flow Rate, Q, by an electromagnetic flow-meter, which
nozzle angles are optimally selected based on the results of CFD was calibrated by the volumetric tank (100 m3) of the test
simulations in the injector. The final dimensions of the injectors stand, with relative uncertainty ±0.5%.

7 Copyright © 2012 by ASME

Downloaded From: http://proceedings.asmedigitalcollection.asme.org/ on 02/01/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use


3) The Relative Static Pressure of the water right before the where M, Q, H and ω are the four quantities measured. Complete
injectors, using a differential pressure transducer (0-20 bar), sets of experiments are performed in almost the entire turbine
with relative uncertainty ±0.25%. operating range that is feasible in this test rig, and with the
4) The Torque, M, developed at the shaft between the turbine upper, the lower or both injectors in operation.
and the brake, by a torque transducer with strain gage
sensing and maximum experimental error ±0.5%.
The Laboratory test rig and the measuring procedure comply 3.5. Results and comparison
with the IEC model test standards. Prior to the experiments all
measuring instruments were calibrated according to their The operation of the Turgo model with the upper or the lower
manuals. nozzle is generally similar, verifying the good quality and
precision of the model manufacturing. The characteristic
operation curves of the overall efficiency are obtained for 15
consecutive opening positions of the nozzles (Fig. 17). Each
such curve is constructed by varying the rotation speed of the
feeding pump and/or the turbine runner. The flow rate results are
presented as function of the dimensionless flow rate parameter
Φ, defined as:
Q
 3
(13)
 Rrun 

and they are converted for comparison to the design speed of the
runner. The spear valve opening fraction, α, is defined as the
distance Δx of the needle from the close valve position, divided
Fig. 15. Completed Turgo model installed in the Laboratory. by the nozzle exit diameter.
The pattern of the efficiency characteristic curves for both
injectors (Fig. 17) is in agreement with the theoretical
performance of impulse hydro turbines: For a given injector
opening there is an optimum flow rate that maximizes the
efficiency. At that point, the remaining angular momentum of
the runner outflow becomes minimum, as confirmed by visual
observation.
The efficiency results like the ones given in Fig. 17 show that
the new Turgo turbine model is capable to attain high
efficiencies for this size, up to 86%, and at a wide range of load
conditions either with a single or with both injectors. In all cases
the envelope curve is quite flat, and the maximum efficiency is
attained for an intermediate spear valve opening.
1.0
Fig. 16. Sketch of the laboratory test rig of Turgo model.
0.9

0.8
The LabView 8.6® graphical programming software is
Turbine Overall Efficiency

0.7
implemented for the data acquisition of the analog signals and
0.6
their conversion into digital. The developed graphical
environment allows for continuous monitoring of the pump and 0.5

turbine operating conditions, and for real time evaluation of a 0.4


measured value, in respect to the turbine performance 0.3 Spear Opening, 
characteristic curves. 0.246
0.369
0.2
The experimental data for the overall turbine efficiency are 0.492
0.738
obtained from the relation: 0.1 0.923

0.0
 0.04 0.06 0.08 0.10 0.12 0.14 0.16 0.18 0.20
 exp  (12) Flow Rate Parameter
 gQH
Fig. 17. Measurements of overall turbine efficiency with the lower
(red) and the upper (blue) injector in operation.

8 Copyright © 2012 by ASME

Downloaded From: http://proceedings.asmedigitalcollection.asme.org/ on 02/01/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Using the developed data base of the experimental results,
complete hill charts of the turbine performance can be The numerical results are compared with the measurements
constructed, as shown in Fig. 18. These charts confirm that the of lower and of upper injectors for various nozzle openings in
best efficiency region of the turbine includes the design point for Fig. 19. The agreement is quite good almost in the entire loading
which it is numerically optimized (324 m3/h, 48 mWG). Also, range of the turbine, and in most operating points the observed
the efficiency remains high for large flow rate variations, while it discrepancies are up to 2-3%. The differences become more
is much more sensitive in head variations, like in a typical significant for the smallest opening (α=0.12, Fig. 11), which
impulse turbine. corresponds to load about 50% of the nominal. This deviation
can be attributed to the assumption of constant mechanical
efficiency degree (97%), instead of constant absolute value of
mechanical losses, which would be more correct for the same
rotation speed. In that case, the percentage portion of mechanical
losses in respect of the shaft power will be double at 50% load
hence the predicted total efficiency would be lower and closer to
the measurements.
On the other hand, the maximum attainable efficiency
reduces as the nozzle opening and the jet diameter increase
significantly above the nominal values (Fig. 19). In this case,
part of the oversized jet flow start to impinge on the hub and
shroud of the runner, and this is also reproduced in the numerical
simulation results.

Fig. 18. Efficiency hillchart of the Turgo model operating with the
upper injector. 4. CONCLUSIONS

The formulation and capabilities of a numerical methodology


The hydraulic and mechanical losses of the runner cannot be developed for flow analysis and design improvement of impulse
directly measured. Consequently, in order to compare the FLS hydraulic turbines are presented and demonstrated in this work.
model output with the measurements, the hydraulic efficiency, The major advantage of the new Fast Lagrangian Simulation
ηh, of Eq. (11) is converted to overall efficiency by the relation: (FLS) algorithm is its fast performance and minimal computer
requirements. This permits its application for multi-parametric
 
  h  2 m (14) and multi-objective design optimization of the runner geometry,
which requires numerous evaluations of the complex, unsteady
flow filed developed during the jet-buckets interaction.
where ηm is the mechanical efficiency of the runner (shaft Parametric and sensitivity analysis of turbine operation or design
bearings and ventilation losses), estimated to 97% for all parameters can be also easily performed.
operating points. Moreover, the FLS model provides the capability to take into
account in an inclusive way the effects of various instabilities
90
and other complex secondary mechanisms evolved during the
jet-bucket interaction, which are difficult or even impossible to
80
be modeled by other simulation approaches.
Total efficiency (%)

70
The main drawback is that the model contains a number of
adjustable coefficients, which must be tuned with the aid of more
60
accurate data, experimental or numerical. However, the present
Lines = Numerical
application showed that the appropriate values of those
50 Symbols = Measur. coefficients are similar for different impulse turbines (Pelton and
α = 0.125
Turgo), as also that after their regulation the model can predict
α = 0.246
40 α = 0.431 with remarkable reliability the turbine performance in a broad
α = 0.615 operation range.
α = 0.923
30 The application of the new model for the optimal design of
0 0.04 0.08 0.12 0.16 0.2 the runner of a new Turgo model turbine was successful, and the
Flow rate parameter, Φ achieved performance and efficiency of the prototype in the
Laboratory is found to be quite close to the predictions.
Fig. 19. Comparison between measured and predicted turbine A more elaborate design would require more detailed and
model performance. accurate modeling, and for this reason further development of

9 Copyright © 2012 by ASME

Downloaded From: http://proceedings.asmedigitalcollection.asme.org/ on 02/01/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use


the FLS model aiming to improve the expressions of the [13] Koukouvinis P., Anagnostopoulos J. and Papantonis D.,
additional terms is currently under way. “SPH Method used for Flow Predictions at a Turgo Impulse
Turbine: Comparison with Fluent”, World Academy of
Science, Engineering and Technology, Vol. 79, 2001, pp.
659-666.
ACKNOWLEDGMENTS [14] Koukouvinis P., Anagnostopoulos J., and Papantonis D.,
“Flow modelling in a Turgo turbine using SPH”, 5th Spheric
This work was carried out in the frame of HYDROACTION
workshop, Manchester, UK, June 23-25, 2010.
Project funded by the European Union (FP7-ENERGY-007-1-
[15] Anagnostopoulos J. and Papantonis D., “A numerical
RTD, Project number 211983).
methodology for design optimization of Pelton turbine
runners,” HYDRO 2006, Porto Carras, Greece, September
25-27, 2006.
[16] Anagnostopoulos J. and Papantonis D., “Flow modeling
REFERENCES
and runner design optimization in Turgo water turbines”,
Intl. Journal of Applied Science, Engineering and
[1] Wilson P.N., “A high-speed Impulse Turbine”, Water
Technology, Vol. 4(3), 2007, pp. 136-141.
Power, January 1967, pp. 25-29.
[17] Nechleba M., “Hydraulic Turbines. Their Design and
[2] Harvey A., “Micro-Hydro Design Manual”, Intermediate
Equipment”. ARTIA, Prague, 1957.
Technology Publications, UK, 2006.
[18] http://velos0.ltt.mech.ntua.gr/research/easy.html
[3] Gibson A.H., “Hydraulics and its applications”. Constable
[19] Giannakoglou K., “Design of optimal aerodynamic shapes
& Co Ltd, UK, 1952.
using stochastic optimization methods and computational
[4] http://mve.energetika.cz/.
intelligence,” Progress in Aerospace Science, Vol. 38,
[5] Staubli T., and Hauser H.P., “A Diagnosis Tool for Pelton
2002, pp. 43-76.
Turbines”. In: IGHEM 2004, 5th Intl. Group for Hydraulic
Efficiency Measurement Conference, Lucerne, Switzerland,
July 14-16, 2004.
[6] Zhang Zh. and Parkinson E., “LDA Application and the
Dual-Measurement-Method in Experimental Investigations
of the Free Surface Jet at a Model Nozzle of a Pelton
Turbine”. In: 11th Intl. Symposium on Applications of Laser
Anemometry to Fluid Mechanics, Lisbon, Portugal, 2002.
[7] Staubli T., Abottspon A., Weibel P., Bissel C., Parkinson E.,
Leduc J., and Leboeuf F., “Jet quality and Pelton
efficiency”. In: HYDRO 2009, Lyon, France, Paper 2.05,
October 26-28, 2009.
[8] Perrig A., Farhat M., Avellan F., Parkinson E., Garcin H.,
Bissel C., Valle M., and Favre J., “Numerical flow analysis
in a Pelton turbine bucket”, Proceedings, 22nd IAHR
Symposium on Hydraulic Machinery and Systems,
Stockholm, Sweden, June 29 – July 2, 2004.
[9] Perrig A., Avellan F., Kueny J.-L., Farhat M., and Parkinson
E., “Flow in a Pelton turbine bucket: Numerical and
experimental investigations”, ASME Trans., Journal of
Fluids Engineering, Vol. 128, 2006, pp. 350-358.
[10] Zoppe B., Pellone C., Maitre T., and Leroy P., “Flow
Analysis Inside a Pelton Turbine Bucket” ASME Trans., J.
of Turbomachinery, Vol. 128, 2006, pp. 500-511.
[11] Santolin A., Cavazzini G., Ardizzon G., and Pavesi G.,
“Numerical Investigation of the Interaction Between Jet and
Bucket in a Pelton Turbine”. Proc. IMechE, Part A: J.
Power and Energy, Vol. 223, 2009 pp. 721-728.
[12] Marongiu J.-Ch., Leboeuf F., and Caro J., Parkinson E.,
“Free surface flows simulations in Pelton turbines using an
hybrid SPH-ALE method”. Journal of Hydraulic Research ,
Vol. 48, 2010, pp. 40-49.

10 Copyright © 2012 by ASME

DownloadedViewFrom:
publicationhttp://proceedings.asmedigitalcollection.asme.org/
stats on 02/01/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use

You might also like