Vous êtes sur la page 1sur 15

Proceedings of the Institution of Mechanical

Engineers, Part A: http://pia.sagepub.com/


Journal of Power and Energy

Conceptual optimization of axial-flow hydraulic turbines with non-free vortex design


R. B F. Albuquerque, N Manzanares-Filho and W Oliveira
Proceedings of the Institution of Mechanical Engineers, Part A: Journal of Power and Energy 2007 221: 713
DOI: 10.1243/09576509JPE394
The online version of this article can be found at:
http://pia.sagepub.com/content/221/5/713

Published by:
http://www.sagepublications.com

On behalf of:

Institution of Mechanical Engineers

Additional services and information for Proceedings of the Institution of Mechanical Engineers, Part A: Journal of Power and Energy can
be found at:
Email Alerts: http://pia.sagepub.com/cgi/alerts
Subscriptions: http://pia.sagepub.com/subscriptions
Reprints: http://www.sagepub.com/journalsReprints.nav
Permissions: http://www.sagepub.com/journalsPermissions.nav
Citations: http://pia.sagepub.com/content/221/5/713.refs.html

Downloaded from pia.sagepub.com by Arik Aprilliyanto on July 22, 2011

713

Conceptual optimization of axial-flow hydraulic turbines


with non-free vortex design
R B F Albuquerque , N Manzanares-Filho, and W Oliveira
Mechanical Engineering Institute, Federal University of Itajub UNIFEI, Itajub, Minas Gerais, Brazil
The manuscript was received on 17 November 2006 and was accepted after revision for publication on 28 March 2007.
DOI: 10.1243/09576509JPE394

Abstract: This paper presents a low cost computational methodology for conceptual design
optimization of axial-flow hydraulic turbines. The flow model away from the blade rows is considered axisymmetric, steady, and with cylindrical stream surfaces. The flow at the cross-sections
behind the distributor and behind the runner is treated by means of the simplified radial equilibrium equation. The flow losses and deviations are assessed by using empirical correlations.
Although simplified, the model allows the consideration of non-free vortex analysis at an early
design stage. For reducing the set of design variables to be optimized, the runner blading stagger,
camber, and chord-pitch ratio are parameterized in terms of their values at the hub, mean,
and tip stations. The optimization problem consists in finding a basic geometry that maximizes the turbine efficiency, given the design flowrate, rotational speed and bounds for the
design variables and also for the available head. Two optimization techniques have been applied:
a standard sequential quadratic programming and a controlled random search algorithm. An
application example is presented and discussed for the optimization of a real turbine model. The
optimal solution is compared with the original turbine design, showing potential performance
improvements.
Keywords: axial-flow hydraulic turbine, non-free vortex, loss and deviation modelling, geometry
parameterization, optimal design

INTRODUCTION

The development of computers in the second half of


20th century made possible the use of complex numerical flow simulation techniques for turbomachine
analysis and design. Nowadays, three-dimensional
Euler codes and three-dimensional viscous Navier
Stokes codes are already standard tools in the development of new turbomachinery units. Details of flow
separation, loss sources, loss distribution in components, matching of components at design and offdesign, and low pressure levels with risk of cavitation
are now amenable to analysis with computational fluid
dynamics (CFD) [1].

Corresponding author: Mechanical Engineering Institute, Federal

University of Itajub, Av. BPS 1303, CP 50, Itajub, Minas Gerais,


37500-903, Brazil. email: galegosigma@yahoo.com.br
JPE394 IMechE 2007

Although three-dimensional NavierStokes codes


have allowed good performance predictions and
contributed for decreasing the costs of turbomachine
model tests, a considerable computational effort has
still to be spent with grid generation and with the solution of the flow governing equations in each numerical
investigation. This issue is even more important in
the case of design optimization: when a geometrical change is made during the optimization process,
complex meshes must be recalculated and the flow
solver must be run again. This high effort prevents
the incorporation of sophisticated NavierStokes simulations throughout the whole design procedure [2].
Actually, the analysis and design of turbomachines
still require the use of simpler methodologies mainly
in preliminary design phases, when the geometry
is not yet completely defined. One example of a
very simplified methodology is the mean streamline
analysis for conceptual optimization of mixed-flow
pumps [3]. For axial flow gas turbines, it is common
Proc. IMechE Vol. 221 Part A: J. Power and Energy

Downloaded from pia.sagepub.com by Arik Aprilliyanto on July 22, 2011

714

R B F Albuquerque, N Manzanares-Filho, and W Oliveira

the use of the simplified radial equilibrium equation or


streamline curvature methods for evaluating the radial
flow variations [4].
For hydraulic turbines, however, the description
of this kind of intermediate methodology is apparently scarce in the open literature. On one hand, one
can found theoretical analysis of the overall performance, which does not account for the effects of
runner blade geometry changes in the flow field [5].
On the other hand, it is not difficult to find modern
works on design optimization using direct CFD analysis, without previous systematic investigations about
favourable geometrical configurations [6, 7].
Therefore, it seems desirable to make available intermediate tools for analysis and design of
water turbines. These tools should furnish a reliable conceptual design, with a simplified but representative geometry for runners and stators and
also favourable trends towards the optimal flow
field. In the present work, one such methodology
is proposed for axial-flow hydraulic turbines. The
application of these low head turbines is expanding worldwide because of the progressive exhaust
of hydropower resources of high and moderate
heads.
The proposed methodology is intended to a conceptual design optimization. It couples a fast flow
solver and optimization techniques. The flow solver
is based on the simplified radial equilibrium equation
and on empirical correlations for flow losses and deviations. Besides allowing fast comparative evaluations
of preliminary configurations, the methodology can
deal with non-free vortex analysis at an early design
stage. Thus, the use of high computational cost tools,
like CFD codes, can be postponed to very final design
phases. In this way, the total design effort could be
substantially reduced.
The design optimization problem is stated early
in section 2. In section 3, the design variables
and the blade geometry parameterization are presented. The flow model for the axial turbine is developed in section 4. A brief explanation about the
selected optimization techniques is given in section
5. In section 6, the methodology is applied to a
small tube type propeller turbine and the optimized
results are compared with a previous design for
that turbine. Concluding remarks are presented in
section 7.

FORMULATION OF THE OPTIMIZATION


PROBLEM

order to maximize the turbine efficiency (objective


function), given the design rotational speed and volumetric flowrate (design point optimization only).
The available head should lie within upper and lower
limits, these being the non-linear constraints of the
problem. There are also lateral constraints for the
design variables, defining the problem feasible region
(design space).
Formally, this can be stated as a constrained minimization problem as follows
minimize f (x)
subject to gi (x)  0, i = 1, . . . , m
xS
where x is the n-dimensional vector of design variables xj ( j = 1, . . . , n). These variables are geometrical
parameters, defined in section 3. The search region S is
defined by upper and lower bounds, xjU and xjL , respectively, for each coordinate of x : S = {x n : xjL 
xj  xjU , j = 1, . . . , n}. The objective function is f (x) =
(x), where is the turbine efficiency. gi (x), i =
1, . . . , m, are the m = 2 constraint functions, namely,
g1 (x) = HL H (x) and g2 (x) = H (x) HU , where H is
the turbine available head and HL and HU are, respectively, lower and upper limits, such that HL  H  HU .
The performance of the turbine ( and H ) is calculated through an appropriate flow solver. The one to
be applied in this work is developed in section 4.
The optimization problem stated above can be
solved by algorithms that treat directly non-linear constraints. Otherwise, the available head constraints can
be imposed by means of a penalization scheme on the
objective function

+ M (HL H )2 , H < HL

f = , HL  H  HU

+ M (H HU )2 , H > HU
where M is a sufficiently large positive number. Again,
the objective is to maximize (minimize ) with
HL  H  HU . The choice of the penalty factor M must
not drive the optimization process towards a penalty
minimization only, missing objective function main
information, i.e.the efficiency . Also, the constraints
must not be violated at the end of the process. Some
tests have to be performed to settle suitable values for
M in each particular problem.
3

The hydraulic turbine design problem consists in


searching some basic geometrical parameters (design
variables) of the guide vanes and runner blades in

BLADE GEOMETRY PARAMETERIZATION

The water turbine considered in this study is a tube


type propeller turbine with non-adjustable guide

Proc. IMechE Vol. 221 Part A: J. Power and Energy

Downloaded from pia.sagepub.com by Arik Aprilliyanto on July 22, 2011

JPE394 IMechE 2007

Conceptual optimization of axial-flow hydraulic turbines

Fig. 1

Sketch of the propeller turbine water channel

vanes, shown in Fig. 1. The distributor is cylindrical


(non-conical) and presents untwisted guide vanes of
constant chord. Thus, only the runner blade geometry
will be parameterized in this work. The parameterization should lead to a small number of design variables without missing the relevant geometric information.
The blade profile camber lines are approximated
by arcs of circumference (ARC) of small curvature, an
acceptable assumption for slightly cambered profiles
as those employed in axial hydraulic turbines [6, 8].
The blade thickness is not considered in this work,
since no cavitation phenomena or flow separation is
addressed by the adopted modelling. Actually, when
the blade profiles of a given cascade are thin enough,
the thickness does not contribute to the flow turning
angle, which becomes a function just of the angle of
attack and profile camber [9]. Thus the runner cascade
geometry is defined by its stagger angle, , chord-pitch
ratio, /t, and relative camber at midchord, f / (Fig. 2).
This choice of design parameters is suitable for computing the relevant kinematics characteristics at any

715

radial station (as incidence angles, angles of attack,


deviation angles, and flow turning angles).
The runner blading stagger, camber and chordpitch ratio spanwise variations are parameterized
in terms of their values at the hub, mean, and
tip stations. In this way, the number of runner
design variables is reduced to nine. Here, parabolic
functions of the radius are chosen for parameterization. This choice can satisfactorily approximate
the usual geometry found in axial hydraulic turbines [7]. Since the guide vanes are untwisted along
the radius, a single outlet angle, 2 , is enough as
design variable for the distributor (Fig. 2). Thus, one
has a total of ten design variables. These variables
have been chosen in order to easily identify performance improvements at an intermediate design
phase.
The application example to be presented in this
study uses as geometrical reference the tube type
propeller turbine designed and tested by Souza [10].
Table 1 shows some basic features of this turbine.
The distributor is untwisted and with constant chord
length. The relevant geometric parameters of the
runner are reasonably reproduced by the parameterization here proposed, as shown in Fig. 3. This
parameterized approximation of the previous geometry [10] is referred here as the original design and will
be useful for comparisons with optimization results
(section 6). Remark: the original design did not employ
ARC blades, but Gttingen profiles. Then, equivalent
ARC blades were calculated for the original design so
that the camber angles of the Gttingen profiles have
been recovered.
It must be emphasized that a conceptual design
is book for here, in which the attained flow velocity
distribution patterns are more relevant than the geometry itself. Naturally, the chosen geometric parameters
should be sufficiently representative for the flow variations imparted by the blade rows. They should also
be relevant for the deviation and loss correlations to
be used in the flow model (section 4). The parameterization here proposed satisfies these requirements.
Further geometrical refinements could be achieved by
using more evolved methodologies out of the scope of
the present work.

Table 1 Turbine main features [10]

Fig. 2 The design parameters in a radial station: (a)


distributor cascade: outlet angle, 2 ; (b) runner
cascade: stagger angle, , chord length, , camber
at midchord, f
JPE394 IMechE 2007

Flowrate
Rotational speed
Head
Efficiency
Output
External diameter
Internal diameter
Number of blades
Guide vanes exit angle

0.267 m3 /s
1145 r/min
3.9 m
82%
8.3 kW
280 mm
112 mm
4
60 (from tangential)

Proc. IMechE Vol. 221 Part A: J. Power and Energy

Downloaded from pia.sagepub.com by Arik Aprilliyanto on July 22, 2011

716

R B F Albuquerque, N Manzanares-Filho, and W Oliveira

Fig. 3

Parabolic parameterizations for the runner geometry of the original design

FLOW MODEL FOR AN AXIAL-FLOW


HYDRAULIC TURBINE

4.1 The simplified radial equilibrium equation


with energy balances through blade rows
It will be considered the assumptions of axisymmetric steady incompressible flow in cylindrical stream
surfaces, such that the simplified radial equilibrium
equation away from the blade rows becomes the only
relevant differential equation for the fluid motion. This
equation relates the meridional (axial) and circumferential velocity components, cm and cu , as functions just
of the radius r and can be stated as follows (11, p. 266,
equation (4.8))
dhS
ds
cu d
dcm
T
=
(rcu ) + cm
dr
dr
r dr
dr

(1)

where hS represents the stagnation enthalpy (hS = h +


c 2 /2), T the absolute temperature and s the entropy.
Using the Gibbs relation dh = dp/ + T ds and the
incompressibility assumption, equation (1) can be
written as
1 dpS
dcm
cu d
=
(rcu ) + cm
dr
r dr
dr

(2)

where pS represents the stagnation pressure (pS = p +


c 2 /2) and is the fluid density.
Besides reference [11], the deduction of the simplified radial equilibrium can be found in many other
textbooks [12, 13]. However, the differences between
the flow conditions behind a stator (where there is no
work transfer) and behind a runner or impeller (where
a spanwise distribution of work transfer occurs) are
not clearly pointed out in those references. Commonly,
the actual energy balance in the upstream blade rows
is not explicitly accounted for.
In what follows, suitable formulations of the radial
equilibrium equation (2) will be presented for the flow

behind the distributor and behind the runner. The


basic procedure consists in applying integral energy
balances through the corresponding upstream blade
cascades and introducing the results in equation (2).
It will be assumed that the stream surfaces remain
cylindrical along the blade rows. It is possible to show
that radial deviations of the stream surfaces associated with changes in meridional (axial) velocities are
indeed negligible in design situations [14]. Iterative
schemes for numerically obtaining the velocity distributions are also proposed. The developments are
referred to in Figs 4 to 7. See also the notation for the
velocities, angles, and subscripts designation.
4.1.1

Radial equilibrium for the flow behind the


distributor

The integral energy equation applied to a distributor


cascade at a certain radius r leads to
pS1 pS2 = YLs

(3)

where YLs is the mechanical energy loss per unit mass


in the distributor cascade.
The stagnation pressure at the distributor inlet, pS1 ,
is assumed to be constant along the span. Therefore, differentiating equation (3) and substituting in

Fig. 4

Meridional cross-section of turbine water channel. 1: distributor inlet; 2: distributor outlet;


4: runner inlet; 5: runner outlet

Proc. IMechE Vol. 221 Part A: J. Power and Energy

Downloaded from pia.sagepub.com by Arik Aprilliyanto on July 22, 2011

JPE394 IMechE 2007

Conceptual optimization of axial-flow hydraulic turbines

717

vane geometry does not furnish the free vortex at


the enclosure between the distributor and the runner, cm2 will not be uniform along the span; thus,
the correct radial equilibrium must be evaluated in
order to achieve a realistic velocity profile prediction.
Moreover, the free vortex condition is not necessarily
the optimal one and so a non-free vortex analysis is
desirable in any automatic design system.
The integration of equation (4) leads to the distribution of cm2 in terms of the velocity torque (rcu2 )

Fig. 5

Instantaneous absolute
cylindrical section

streamlines

in

2
2
cm2
(r) cm2h
= 2[YLs (rh ) YLs (r)]
r
1 d(rcu2 )2
+
dr = Is (r)
2
r
dr
rh

(5)

or
cm2 (r) =

Fig. 6 Velocity components at distributor cascade


2
cm2h
+ Is (r)

(6)

As a first approximation, the radial variations of


energy loss will be considered negligible in comparison with the radial variations of meridional kinetic
energy in equation (5). However, it must be emphasized that the losses will be computed for evaluating
the turbine efficiency, after the flow calculation.
The overall continuity must be imposed in order
to evaluate the meridional velocity at the hub station (cm2h ). Representing the elementary volumetric
flowrate by dQ = 2rcm dr, integrating between hub
and tip and using equation (6), one obtains
 rt 
2
cm2h
+ Is (r)r dr = Q/2
(7)
rh

Fig. 7 Velocity triangles at runner cascade

equation (2) with subscript 2, one obtains an expression of the radial equilibrium equation suitable for the
flow calculation behind the distributor
cu2

d(rcu2 )
dcm2
dYLs
+ rcm2
+r
=0
dr
dr
dr

(4)

Observe in equation (4) that if one neglects the


radial loss variation (dYLs /dr = 0), the free vortex condition rcu2 = const. implies that cm2 = const. and viceversa. The free vortex is a classical design approach
in axial hydraulic turbines. However, if the guide
JPE394 IMechE 2007

The distributions cm2 (r) and cu2 (r) will be determined by solving equations (6) and (7) iteratively.
In any iteration, equation (7) is treated as a nonlinear algebraic equation for the unknown cm2h with
a given and fixed flowrate. A standard bisection-based
algorithm was chosen to solve this equation in conjunction with the Simpson rule for the necessary
integral evaluations.
The evaluation of Is (r) in equation (5) requires the
previous knowledge of the velocity torque distribution, rcu2 (r). However, the circumferential component
cu2 at a certain cascade depends upon the not yet
determined meridional component cm2 . Therefore, an
iterative scheme is adopted. One first assumes a uniform distribution to cm2 . Thence, some rcu2 values
are calculated by using cascade relations in N radial
stations. These values are fitted to a parabolic distribution, rcu2 = K1 + K2 r + K3 r 2 , by using least-squares.
This choice is indeed suitable for axial hydraulic turbines, since it is able to satisfactorily reproduce typical
swirl patterns behind the distributor [15] and also
includes the free vortex as a particular case. Thence,
the fitted parabolic distribution is used for evaluating
Proc. IMechE Vol. 221 Part A: J. Power and Energy

Downloaded from pia.sagepub.com by Arik Aprilliyanto on July 22, 2011

718

R B F Albuquerque, N Manzanares-Filho, and W Oliveira

Is (r) analytically in equation (5) and the meridional


velocity distribution in equation (6) can be evaluated
after solving equation (7) for cm2h . This new cm2 distribution is now used to recalculate the rcu2 values in the
N cascades and the iterations are carried out until the
flow field behind the distributor converges (Fig. 8).
4.1.2

 rt 

Radial equilibrium for the flow behind the


runner

rh

The flow field at the runner inlet is assumed to be same


as that at the distributor outlet, which has just been
evaluated. The integral energy equation applied to a
runner cascade at a certain radial coordinate r leads
to
pS4 = pS5 + YLr + Yblade

(8)

where YLr is the mechanical energy loss per unit


mass and Yblade is the blade specific work of the runner cascade, calculated according to the Euler work
equation
Yblade = u(cu4 cu5 )

(9)

Substituting equation (9) into equation (8), differentiating and considering the result in equation (2) with
subscript 5, one obtains an expression of the radial
equilibrium equation suitable for the flow behind the
runner
(cu5 u)

dcm5
d(rcu5 )
dYLr
d(ucu4 )
+ rcm5
+r
+r
=0
dr
dr
dr
dr
(10)

Observe that the distribution of uc u4 is the same as


that of uc u2 already calculated. Again, the integration
of equation (10) leads to the distribution of cm5 in terms
of the distribution of cu5
2
2
cm5
(r) cm5h
= 2[YLr (rh ) YLr (r)]
r
(cu5 u) d(rcu5 )
2
dr
r
dr
rh


2 ucu4 |r ucu4 |rh = Ir (r)

Fig. 8

As a first approximation, the loss variations in


equation (11) were also neglected.
Again, the overall flow continuity is imposed for
evaluating the meridional velocity at the hub station
(cm5h ), leading to a non-linear problem analogous to
that one for the distributor

(11)

Iterative scheme for the velocity distribution evaluation at the exit sections

2
cm5h
+ Ir (r)r dr = Q/2

(12)

Due to the same considerations made for the flow


behind the distributor, an analogous iterative scheme
is adopted for the evaluation of the velocity distribution behind the runner (Fig. 8). Using least-squares
again, the values of rc u5 in N cascades are now fitted to a cubic function (rcu5 = K4 + K5 r + K6 r 2 + K7 r 3 )
instead of a parabolic function. This choice has proved
to be suitable for reproducing with sufficient accuracy
the typical inflections that may occur in cu5 spanwise
variation. For the runner, however, a subrelaxation
scheme had to be applied for attaining convergence.
Each time a new distribution of cm5 is calculated by
equations (12) and (11), leading to corresponding cu5
values in the N cascades (from velocity triangles and
deviation correlation), the new values settled to cu5 are
given by
new
cascade
old
= cu5
+ (1 )cu5
cu5

(13)

where is the subrelaxation factor. For starting this


scheme, the first cu5 distribution is equated to zero.
The subrelaxation factor has been settled equal to 0.10
for a satisfactory convergence rate.
4.2

Loss and deviation correlations; turbine


efficiency

The empirical loss correlations used in this study to


calculate the losses through an axial hydraulic turbine are summarized in Table 2. The recommended
range for the empirical coefficients and the adopted
values are also indicated. This set of loss models was
chosen with the aim of covering the main sources
of loss in a tube type propeller turbine. For conceptual optimization purposes the level of accuracy of
the loss predictions can be considered secondary. The
key point is the capability of the loss modelling in
indicating the correct trends of loss variations due
to geometrical changes, so that different designs can
be judged in a comparative sense [16, 17]. The set
of loss models should drive the optimization search
towards a geometrical configuration with favourable
hydrodynamic characteristics.
The flow deviation with regard to the outlet geometrical angle in any cascade is assessed by the correlation
of Carter and Hughes [12]. The absolute flow deviation

Proc. IMechE Vol. 221 Part A: J. Power and Energy

Downloaded from pia.sagepub.com by Arik Aprilliyanto on July 22, 2011

JPE394 IMechE 2007

Conceptual optimization of axial-flow hydraulic turbines

719

Table 2 The set of loss models used for axial hydraulic turbines
Loss mechanism

Loss model

Reference

Guide vane profile loss (skin


friction loss at stator)
Incidence loss (shock loss at
runner inlet)

YLs = s c22 /2

[12]

2 /2
YLinc = winc

[18]

= 0.5 to 0.7 (set value = 0.5)

cm4
cm4
+
u
winc =
tan 4blade
tan 2f
2
YLr = r w5 /2

[12]

Runner blade profile loss


(skin friction loss at runner)
Draft tube loss (diffusion and
swirl losses)

YLdt = XDm

2
cm5
c2
+ XDu u5
2
2

[8]

XDm = 0.09 or 0.12; XDu = 0.20 to 0.40


(set value = 0.09) (set value = 0.40)
mec = 0.95 to 0.99 (set value = 0.95)

Mechanical loss (external


loss)
Coefficients for the profile loss (correlation of Soderberg):
= (105 /Re )1/4 [(1 + 1 )(0.975 + 0.075b/B) 1]
1 = 0 e 0.01053 , 0 = 0.04 to 0.06 (set value = 0.04)
= 1f 2f or = 4f 5f (flow turning angle)
B/b = radial/axial blade lengths (cascade aspect ratio)
Re = VDh /, V = c2 or w5 , = dynamic viscosity
Dh = 2Bt cos 2f /(t cos 2f + B) or
Dh = 2Bt cos 5f /(t cos 5f + B)

angle at outlet of a distributor cascade, , is given by



= 2f 2vane = m t/

(14)

where 2f is the flow angle, 2vane is the guide vane


geometrical angle, is the profile camber angle ( =
1vane d2vane ),  is the profile chord, t is the spacing, and m is an empirical factor. In reference [12],
m is graphically provided as a function of the stagger angle, , and the kind of camber line (circular or
parabolic). This graph was approximated by the linear
function m() = 0.21 0.04(90 )/60, for circular
camber lines adopted in this work. The relative flow
deviation at a runner cascade exit is similarly evaluated
using the corresponding runner cascade geometry.
In the course of the investigation, it was verified that
the original correlation, equation (14), underpredicted
the deviation angles for slightly cambered blade profiles as those used in runners of axial water turbines.
The correlation of Carter and Hughes was developed
from low speed tests of gas turbine cascades with high
cambered blades. Then, a simple overprediction based
on the original correlation has been used

= 1.5m t/

[12]

this work, further work seems still necessary to develop


reliable empirical correlations for use in hydraulic
turbine design.
The losses are evaluated in N radial stations (cascades) and these values are adjusted to a cubic
polynomial by using least-squares. Thence, these
regressions are integrated along the span in order
to evaluate the total hydraulic power loss, PL . The
runner blade specific work in equation (9) is also
integrated along the span for obtaining the total
blade power, Pblade . From these results, one calculates the available head, H = (Pblade + PL )/(gQ), the

(15)

This overprediction was calibrated in order to reproduce the available head of the turbine tests reported
in reference [10]. For uniformity reasons, it was also
applied for the distributor cascades.
The formula in equation (15) and the loss correlations should be used with some care. Although they are
suitable for the conceptual optimization purposes of
JPE394 IMechE 2007

Adopted

Fig. 9

Overall scheme of the solver code

Proc. IMechE Vol. 221 Part A: J. Power and Energy

Downloaded from pia.sagepub.com by Arik Aprilliyanto on July 22, 2011

720

R B F Albuquerque, N Manzanares-Filho, and W Oliveira

hydraulic efficiency, h = Pblade /(Pblade + PL ) and the


overall efficiency, = h mec . The integrals mentioned
in this paragraph are carried out in a mass flux basis.
The flow chart of the solver described in this section is
shown in Fig. 9. It was coded in MatLabTM language.
As described in section 2, the values of efficiency
(objective function) and available head (non-linear
constraints) are required for solving the proposed
optimization problem.
4.3

Some illustrative results using the flow solver

In Fig. 10, a comparison is made between the free


vortex hypothesis and the results obtained with the
application of the flow solver just described to the original design of Souza [10]. Since the guide vanes are
untwisted along the span, the free vortex is not produced at the enclosure between the distributor and the
runner, thus cm2 is not uniform (Fig. 10(a)). As shown
in Fig. 10(b), the blade work transfer varies from hub
to tip, although the average value is in good agreement
with the free vortex hypothesis. These results emphasize that considering the correct radial equilibrium is
indeed important for good predictions of the velocity
profiles downstream and upstream the runner. As the
runner blades were not specifically designed for satisfying these profiles, it seems clear that there is room
for design improvements. Moreover, the velocity profiles themselves may not be the most favourable ones
for the given operational conditions (flowrate, rotational speed, and available head). In fact, these issues

have motivated the development of the proposed conceptual design optimization tool using the flow solver
described in this section.

5 THE CHOSEN OPTIMIZATION METHODS


The conceptual optimization methodology is completed by coupling the flow solver with a suitable
optimization technique. Two optimization methods
have been alternatively tested for this aim: a sequential
quadratic programming method (SQP) and a controlled random search algorithm (CRSA). The SQP is
a gradient based method useful for local searches
starting from a previous design. The CRSA is a population set-based direct search algorithm that helps
in exploratory searching throughout the whole design
space.
The SQP method is one of the most efficient optimization techniques for solving constrained non-linear problems being suitable for the
present application [19]. The fmincon function from
MatLabTM was chosen in this work. This is an efficient
implementation of the standard SQP using the BFGS
formula for approximating the Hessian matrix [19]. In
the present application, the option for evaluating the
directional derivatives by finite differences was set.
The two main limitations of gradient based methods
are the search for local optimizers only and the need
of a starting guess for the design variables. The success

Fig. 10 Velocity profiles (a) and blade work transfer (b) in the original design
Proc. IMechE Vol. 221 Part A: J. Power and Energy

Downloaded from pia.sagepub.com by Arik Aprilliyanto on July 22, 2011

JPE394 IMechE 2007

Conceptual optimization of axial-flow hydraulic turbines

of the search becomes very dependent of this starting


guess and requires that the designer provide an initial
configuration not too far from an acceptable optimum. Moreover, previous investigations have shown
that even slightly different starting points (designs)
can lead to distinct solutions with different values of
efficiency [20].
To try overcoming these limitations, the CRSA has
also been applied. The CRSA was first proposed by
Price [21] and substantially improved by Ali et al. [22].
Like genetic and differential evolution algorithms, the
CRSA is a global population set-based algorithm. It
starts with an initial population of points on the design
space and then performs iterative substitutions of
worst points by better points in order to contract the
whole population towards a global optimizer. In CRSA,
a single point is replaced per iteration. The CRSA was
chosen because of its straightforward implementation, fastness, and good results reported in technical
literature [23]. Here one applies the algorithm proposed by Ali et al. [22] with modifications for avoiding ill-conditioning and accelerating the convergence
when solutions lie in the vicinity of the design space
boundaries [20].
Differently of the SQP, the CRSA does not require a
careful starting design. It employs an initial population
randomly chosen on the design space S. Besides to
alleviate the designer effort, the CRSA increases the
hope of finding a global solution. A relatively small
number of function evaluations for convergence is also
an important feature of CRSA [20].
When using the CRSA, the available head constraints
are imposed by means of the penalty scheme already
described in section 2.
6

methodology proposed in this work. The optimization


runs were performed according to Table 3 for the
operating point and head constraints and Table 4
for design variables lateral constraints. The analysis
of the original design and the best results obtained
by optimization are compared in Table 5. The values
marked with an asterisk correspond to an activated
constraint.
Several starting points were tried for the SQP
method. Different solutions have been found, what
denotes the existence of local minima in this problem.
The SQP runs lead to values for the guide vane outlet
angle, 2 , ranging from 51 to 68 . This is a main concern since the distributor exit flow strongly affects the
design of the runner blades and, therefore, the overall turbine performance. The SQP solution shown in
Table 5 is the best one among these configurations.
Curiously, this solution was obtained using the original
design as the starting guess.
The CRSA was run 30 times with different initial populations. A penalty factor M = 50 m2 was employed.
In comparison with the SQP, the obtained guide vane
outlet angles lied in a narrower range (50 to 55 ) what
suggests that the concerning global optimizer is within
this range. The CRSA solution shown in Table 5 is the
best one found. It shall be noted the good agreement
between the SQP and the CRSA solutions. Therefore,
it is hoped that both solutions are close to the actual
global optimizer. Since the best SQP solution presents
a slightly higher efficiency than that of the best CRSA
solution, the first is treated as the optimal design in the
following discussion.
The optimal distributor exit angle, 2 = 52.6 , yields
a most favorable flow distribution at the runner inlet.
In comparison with the original design, this makes
possible a more efficient absorption of the flow energy
by the runner. Figures 11 and 12 compare the spanwise variation of the runner blade geometry between
original and optimal designs. The blades of the optimal design have lower stagger angles near the hub and
higher ones near the tip (Fig. 11(a)). They are more
cambered near the hub and less cambered along the
remaining span (Fig. 11(b)). A sketch of the optimal
runner blade is shown in Fig. 13.
The angular momentum of the optimized runner
inlet flow is higher than that of the original design.
This can be seen in Fig. 14(a), cu2 curve. The optimized runner produces a more uniform exit flow (cm5

APPLICATION EXAMPLE

The original design of Souza [10] is used as a comparative reference for an application of the design

Table 3

Design point and head


constraints
0.267 m3 /s
1145 r/min
3.0 m
4.0 m

Flowrate, Q
Rotational speed, n
Lower head, HL
Upper head, HU

Table 4

721

Design variables lateral constraints

( )

/t ()

f / (%)

Design
variable

Hub

Mean

Tip

Hub

Mean

Tip

Hub

Mean

Tip

2 ( )

xjL

40

25

15

1.61

1.08

0.889

0.8

0.5

0.1

50

xjU

55

35

25

1.70

1.20

1.00

6.0

4.0

2.0

70

JPE394 IMechE 2007

Proc. IMechE Vol. 221 Part A: J. Power and Energy

Downloaded from pia.sagepub.com by Arik Aprilliyanto on July 22, 2011

722

R B F Albuquerque, N Manzanares-Filho, and W Oliveira

Table 5

Comparison between original design, best SQP solution and CRSA solution

Design variables and


resulting quantities
( )
/t ()
f / (%)
2 ( )
Blade power (W)
Distributor loss (W)
Runner + draft tube loss (W)
(%)
H (m)

Fig. 11

Original design [10]

CRSA (present work)

Hub

Mean

Tip

Hub

Mean

Tip

Hub

Mean

Tip

49.3
1.61
4.40

26.2
1.08
3.06
60.0
9414
129
894
85.69
4.00

17.4
0.889
1.32

45.8
1.61
5.24

25.9
1.08
2.35
52.6
9534
151
744
86.84
4.00

18.9
0.889
0.27

49.5
1.64
4.72

26.6
1.082
2.15
52.3
9527
152
745
86.82
4.00

17.3
0.902
1.00

Runner blading (a) stagger angle and (b) camber of original design and optimized solution

curve), decreasing the meridional component of the


draft tube loss. Although the exit swirl is everywhere
negative in the original design, the optimal solution
shows a negative swirl near the hub and a positive one
at the tip (Fig. 14(a), cu5 curve, and Fig. 14(b), 5f curve).

Fig. 12

SQP (present work)

Comparison between original design geometry and optimal one (the thicknesses are only
illustrative)

This trend is in good agreement with flow measurements in well-designed axial hydraulic turbines [24].
The optimized flow turning angles in the runner vary
from 22 at the hub to 1 at the tip (Fig. 14(b)) being
higher than those of the original design from hub

Fig. 13

Optimal runner blade (the thickness is only


illustrative)

Proc. IMechE Vol. 221 Part A: J. Power and Energy

Downloaded from pia.sagepub.com by Arik Aprilliyanto on July 22, 2011

JPE394 IMechE 2007

Conceptual optimization of axial-flow hydraulic turbines

Fig. 14

723

Comparison between original and optimal (a) flow velocity distributions, (b) flow angle
distributions, (c) blade work transfer distribution, and (d) spanwise loss variations

to midspan. However, they still remain below safe


limits attainable by axial-flow hydraulic turbines [5].
This issue is also important for avoiding cavitation
risk, since the blade loading has a direct impact in
cavitation phenomena.
Figure 14(c) shows the spanwise distribution of the
blade specific work. In comparison with the original design, the optimal solution increased the blade
JPE394 IMechE 2007

specific work from hub to midspan and decreased it


from midspan to tip. The resulting overall blade power
(Pblade ) was increased in 1.3 per cent (Table 5).
Figure 14(d) shows the spanwise variation of the
specific losses. The sum of runner and draft tube
losses is also plotted, showing the significant decrease
achieved by the optimization (17 per cent, Table 5).
Note, however, that the distributor loss had to be
Proc. IMechE Vol. 221 Part A: J. Power and Energy

Downloaded from pia.sagepub.com by Arik Aprilliyanto on July 22, 2011

724

R B F Albuquerque, N Manzanares-Filho, and W Oliveira

increased to accomplish the higher flow deflections


produced by the optimal guide vanes. The main
improvements occur in the incidence and draft tube
losses. This occurs because the optimal distributor
and runner geometry together provide favourable
incidence angles at the runner inlet and favourable
velocity profiles for the draft tube inlet flow. This is
a key point: the final design of the runner blades
should be made in order to satisfy the optimal velocity
distributions. More sophisticated design tools could
be later used to reduce the actual losses to minimum levels. Thus, among the results obtained by
the conceptual design, one may attribute more significance to the optimal velocity profiles than to the
geometry itself. Nevertheless, the conceptual optimized geometry can be useful as a preliminary
design.
Finally, it is interesting to make a comment about
the computational effort. The runs were on an AMD
SempromTM 2.0 GHz processor with MatLabTM version 5.3 under Windows XPTM platform. The SQP runs
take 2 min in average, spending from 200 to 600 function evaluations (flow solver calls). The CRSA runs
take 10 min in average (from 600 to 1500 function
evaluations), which can also be considered a reasonable time for a conceptual design. Because of
its stochastic features, any CRSA solution should be
taken in a statistic sense. Actually, one should run
several times the CRSA in order to accept a final
solution. If the flow solver demands high computational effort and the global optimization algorithm
requires too many function evaluations, it would be
prohibitive to perform a comparative study, or even
to perform a single optimization run until achieving a
fair convergence.

CONCLUSIONS

A conceptual design optimization methodology for


axial-flow hydraulic turbines has been proposed.
The simplified radial equilibrium equation plays
a central role in the present work. In conjunction
with energy balances through blade rows, it allows
a consistent non-free vortex flow analysis behind
the distributor and behind the runner. A geometrical parameterization is used to represent the spanwise
variation of the blade geometry, giving a small set of
design variables. Loss and deviation correlations complete the flow model. This methodology, coupled with
numerical optimization techniques, allows the calculation of favourable flow trends at an early design
phase.
The methodology was applied to a preliminary
design of a small tube type propeller turbine. The
analysis of the optimized solution shows potential

performance improvements in comparison with a


previous design.
After applying the proposed design system, a final
design of the guide vanes and runner blades could be
made in order to satisfy the obtained velocity distributions. More sophisticated design tools could be used
to reduce the actual losses to minimum levels. Thus,
among the results obtained by the present design
methodology, one may attribute more significance to
the velocity profiles than to the geometry itself. Nevertheless, the obtained optimal parameters can be useful
for reducing the subsequent design effort.

ACKNOWLEDGEMENT
During this work, the first author received financial
support from CAPESCoordenao de Aperfeioamento de Pessoal de Nvel Superior, Brazilian Government Agency.

REFERENCES
1 Drtina, P. and Sallaberger, M. Hydraulic turbines
basic principles and state-of-the-art computational fluid
dynamics applications. Proc. Instn Mech. Engrs, Part C: J.
Mechanical Engineering Science, 1999, 213, 85102.
2 Hirsch, C. and Demeulenaere, A. State of the art in the
industrial CFD for turbomachinery flows. In A thematic
network for quality and trust in the industrial application
of CFD, QNET-CFD Network Newsletter 2003, vol. 2, pp.
59.
3 Oh, H.W. and Kim, K.-Y. Conceptual design optimization
of mixed-flow pump impellers using mean streamline
analysis. Proc. Instn Mech. Engrs, Part A: J. Power and
Energy, 2001, 215, 133138.
4 Yoon, E. S., Kim, B. N., and Chung, M. K. Modeling of
three dimensional unsteady flow effects in axial flow
turbine rotors. Mech. Res. Commun., 1998, 25, 1524.
5 Parker, G. J. A theoretical study of the performance of
an axial flow turbine for a microhydro installation. Proc.
Instn Mech. Engrs, Part A: J. Power and Energy, 1996, 210,
121129.
6 Kueny, J.-L., Lestriez, R., Helali, A., Demeulenaere, A.,
and Hirsch, C. Optimal design of a small hydraulic
turbine. In Proceedings of 22nd IAHR Symposium on
Hydraulic Machinery and Systems, Stockholm, 2004,
paper A02-2.
7 Lipej, A. Optimization method for the design of axial
hydraulic turbines. Proc. Instn Mech. Engrs, Part A: J.
Power and Energy, 2004, 218, 4350.
8 Raabe, J. Hydro power, 1989 (VDI Verlagembh, Dusseldof).
9 Karamcheti, K. Principles of ideal-fluid aerodynamics,
1980 (Robert E. Krieger Plublishing Company, Florida).
10 Souza, Z. Test and operation reports for the MEP tube
type propeller turbine (in Portuguese). Private report,
UNIFEI, 1989, Itajub, Brazil.

Proc. IMechE Vol. 221 Part A: J. Power and Energy

Downloaded from pia.sagepub.com by Arik Aprilliyanto on July 22, 2011

JPE394 IMechE 2007

Conceptual optimization of axial-flow hydraulic turbines

11 Lakshminarayana, B. Fluid dynamics and heat transfer


of turbomachinery, 1996 (John Wiley & Sons, New York).
12 Horlock, J. H. Axial flow turbines, 1973 (Robert E. Krieger
Publishing Company, Huntington).
13 Dixon, S. L. Fluid mechanics, thermodynamics of turbomachinery, 1998 (Butterworth-Heinemann, Oxford).
14 Downie, R. J.,Thompson, M. C., and Wallis, R. A. An engineering approach to blade designs for low to medium
pressure raise rotor-only axial fans. Exp. Therm. Fluid
Sci., 1993, 6, 376401.
15 Peng, G., Cao, S., Ishizuka, M., and Hayama, S. Design
optimization of axial flow hydraulic turbine runner: part
I an improved Q3D inverse method. Int. J. Numer.
Methods Fluids, 2002, 39, 517531.
16 Denton, J. D. Loss mechanisms in turbomachines. Trans.
ASME, J. Turbomach., 1993, 115, 621656.
17 Casey, M. Best practice advice for CFD in turbomachinery design. In A thematic network for quality and trust in
the industrial application of CFD, QNET-CFD Network
Newsletter, 2003, 2, 3537.
18 Pfleiderer, C. and Petermann, H. Mquinas de fluxo,
1979 (LTC, Rio de Janeiro).
19 Nash, S. G. and Sofer, A. Linear and nonlinear programming, 1996 (McGraw-Hill, New York).
20 Albuquerque, R. B. F. Design of axial-flow hydraulic
turbines with geometry parameterization, the radial equilibrium equation and optimization techniques (in Portuguese). MSc Dissertation, Universidade Federal de
Itajub, Itajub, Brazil, 2006.
21 Price, W. L. A controlled random search procedure for
global optimisation. Comput. J., 1977, 20, 367370.
22 Ali, M. M., Trn, A., and Viitanen, S. A numerical comparison of some modified controlled random search
algorithms. J. Glob. Optim., 1997, 11, 377385.
23 Ali, M. M., Storey, C., and Trn, A. Application of
some stochastic global optimization algorithms to
practical problems. J. Optim. Theory Appl., 1997, 95,
545563.
24 Osterwalder, J. Flow measurements on models as a
means for determining the loss distribution in Kaplan
turbines. Escher Wyss News, 1960, Vol. 1, 313.

APPENDIX
Notation
c
f
f /
g
gi
h
H
K1,...,7

absolute velocity (m/s)


objective function
relative cambering of the profile
acceleration due to gravity (m/s2 )
non-linear constraints
specific enthalpy (J/kg)
turbine available head (m)
coefficients of polynomial fitting

JPE394 IMechE 2007

725

chord-pitch ratio
empirical factor for the deviation
correlation
penalty factor (m2 )
rotational speed (r/min), number
of variables
number of cascades
pressure (Pa)
power (W)
flowrate (m3 /s)
radius (m)
specific entropy (J/K)
search region in n
spacing (m)
absolute temperature (K)
circumferential velocity (m/s)
relative velocity (m/s)
vector of design variables
energy per unit mass (J/kg)

/t
m
M
n
N
p
P
Q
r
s
S
t
T
u
w
x
Y

absolute flow angle, guide vane


outlet angle (from tangential) ( )
relative flow angle, stagger angle
(from tangential) ( )
deviation angle ( )
flow turning angle in a cascade ( )
efficiency
subrelaxation factor, loss coefficient due to incidence
dynamic viscosity (Pa s)
loss coefficient due to skin friction
density (kg/m3 )
profile camber angle ( )

Subscripts
blade
f
h
inc
L
m

absorbed by the runner blades


flow
hub radius, hydraulic
incidence or shock
loss, lower bound
meridional component, mean
radius
mechanic
stagnation
tip radius
circumferential component
upper bound
distributor inlet
distributor outlet
runner inlet
runner outlet

mec
S
t
u
U
1
2
4
5

Proc. IMechE Vol. 221 Part A: J. Power and Energy

Downloaded from pia.sagepub.com by Arik Aprilliyanto on July 22, 2011

Downloaded from pia.sagepub.com by Arik Aprilliyanto on July 22, 2011

Vous aimerez peut-être aussi