Вы находитесь на странице: 1из 9

2nd International Conference “From Scientific Computing to Computational Engineering”

2nd IC-SCCE
Athens, 5-8 July, 2006
© IC-SCCE

PARAMETRIC STUDY AND DESIGN OPTIMIZATION


OF A RADIAL FLOW PUMP IMPELLER

Vasilios A. Grapsas, John S. Anagnostopoulos and Dimitrios E. Papantonis

Laboratory of Hydraulic Turbomachines


School of Mechanical Engineering / Fluids Section
National Technical University of Athens, Greece
Heroon Polytechniou 9, Zografou, 15780 Athens, Greece
e-mail: vgrapsas@fluid.mech.ntua.gr

Keywords: Hydraulic turbine blades design, numerical simulation, shape parameterization, optimization.

Abstract. In the present work a parametric study of a radial flow pump impeller concerning the wrap angle of
the blades is presented. The design scenario contemplated here involves the maximization of impeller’s
efficiency by adjusting the blade’s wrap angle and the number of blades. The method is used to improve the
performance of an impeller of known characteristics. An optimal solution is proposed and compared to the
original configuration. For the numerical simulation, the viscous Navier-Stokes equations are handled with the
control volume approach and the k-ε turbulence model. An unstructured Cartesian grid is applied. Advanced
numerical techniques for adaptive grid refinement and for the partially blocked cells are also implemented to
treat the irregular boundaries of the blade. The numerical optimization is performed by a stochastic
evolutionary algorithm, which helps the designer to achieve the required optimum impeller blade that maximizes
certain flow characteristics, like the efficiency or the H-Q curve. The results show that the developed
methodology can be used for the prediction of the optimum geometry among a number of different impellers.

1 INTRODUCTION
The increasing performance levels and operating conditions requirements make the task of designing a pump
impeller very challenging. The role of internal flows and the viscous effects in centrifugal impeller blades are
fundamental and must be taken properly into account in the design process in order to obtain optimum
performance. Computational fluid dynamics (CFD) analysis is being increasingly applied in the design of
centrifugal pumps and nowadays the complex internal flows in water pump impellers can be well predicted to
speed up the pump design procedure.[1] To be able to generate a large panel of blade geometries, a large set of
geometrical parameters is needed,[2] while empirical coefficients are used to describe the effects of entrance
geometry, incident angle, blade thickness and non-uniformity of the inlet velocity profile.[3] A numerical
analysis with such a number of parameters is time consuming, without insurance to converge to an acceptable
solution. The recent progress of the CFD performance enables to reduce the design time by coupling CFD codes
with optimization tools. Optimization techniques are becoming more and more standard for the design
evaluation and improvement.[4]
The possibility of correlating the hydraulic parameters with the geometry of the pump and calculating the
resulting losses or efficiency allows to examine the variations of performance with geometry.[5] Assuming that
the evaluation of the losses is adequately accurate, a systematic variation of the geometry should reveal the
presence of an optimum in terms of geometrical ratios.
However, the influence of each constraint on the pump’s performance must be considered before an
effective optimization is performed. Of high importance decision variables are the blade’s leading and trailing
edge and some curvature parameters of the blade shape. Since a comprehensive treatment of blade’s leading
edge has already carried out,[6] a detailed study of blade’s curvature effect to the impeller performance is
presented herein.
In the present paper an investigation of the influence of the blade’s length, corresponding to the wrap angle,
was carried out in order to improve the pump performance. The method is based on the combination of a
numerical code which creates a 2D blade geometry and calculates the flow field and an evolutionary
optimization algorithm. The predicted results for the head, efficiency and minimum pressure developed in the
impeller are presented over the entire flow range. The optimum wrap angle in terms of best efficiency is found
using a design optimization algorithm. The latter has been recently tested for laminar flows in simple geometries
[7]
with quite encouraging results as well as for the optimization of the leading edge [6] of the blades of the
impeller examined here. Finally, the above study was repeated for a different number of blades (z=6, 7, 8, 10
Vasilios A. Grapsas, John S. Anagnostopoulos, and Dimitrios E. Papantonis.
and 11) and the number of impeller blades maximizing the hydraulic efficiency is also revealed.

2 MODEL FORMULATION
2.1 Flow solver
For two-dimensional, incompressible, turbulent flow, the continuity and momentum equations can be written
in the rotating cylindrical coordinate system as follows:
G
Continuity: ∇⋅w = 0 (1)
G G G G G G G ∇p G
Momentum: w ⋅ ∇w = −2ω × w − ω × (ω × w) − +ν ∇2w (2)
ρ
G
where p is the fluid pressure, ρ is the density, ν is the kinematic viscosity, w is the fluid velocity in the rotating
system and ω is the rotation speed of the impeller. In the calculations, the system of the above equations is
numerically solved with the finite-volume method whereas turbulence is modelled with the standard k-ε model.
An unstructured grid system is used for the solution of the discretized governing equations, [8] which is
generated in a fully automated way and is equipped with adaptive refinement and partially blocked cell
techniques. [9,10] A pseudo-3D incompressible term in dimensional form is added to the continuity equation to
simulate the actual velocity and pressure field in the 3D blade-to-blade domain of the impeller.
The main drawback of the Cartesian grid, which is the simulation of irregular boundaries came over with a
“partially blocked cell” technique, [11] which makes possible the solution of the grid cells that are crossed by the
blade surfaces with an almost second order of accuracy. During the optimization process the evaluation
algorithm is capable to automatically generate the grid and solve the flow equations for any blade geometry that
is generated according to the values given by the optimizer. For stability reasons, the algorithm starts from low
rotational speed and gradually reaches the nominal speed of 3000rpm.

2.2 Geometry Representation


The blades in a centrifugal impeller must guide the flow in a way to minimize the hydraulic losses through
the impeller. The objective when designing an impeller’s blade is to find the best geometry for a certain
operating range. A basic parameter is the blade’s length which can be different for the same inner and outer
diameters D1 and D2, for the same angles β1 and β2 and for the same number of blades. This length plays an
important role in the blade design process, since modification of the blade length results in modifications to the
passage width between two consecutive blades. The blades profile geometry can be described in a simple and
flexible manner using the point by point method.[12, 13] This method of determining blade shapes, introduced by
C. Pfleiderer, is based on the assumption that the transition from β1 to β2 and from r1 to r2 depends on the radius
r. After determining the first point (β1, r1), the last point (β2, r2) and an intermediate
( RBM ⋅ ( β1 + β 2 ) , (r1 + r2 ) ), a smooth curve is drawn through them, giving the mean camber line of the blade.
2 2
The local blade angle β (Fig. 1) is related with the circumferential angle dθ, in polar coordinates as
dr
tgβ (r ) = (3)
r ⋅ dθ
where β(r) is the blade angle distribution along the radius r, which follows a second order polynomial defined as
:
β (r ) = a ⋅ r 2 + b ⋅ r + c (4)
where
(β + β 2 ) ⋅ (RBM − 1) (5)
a=− 1
(r2 − r1 )2
(
b = β 2 − β 1 − a ⋅ r22 − r12 ) (6)
(
c = β 1 − a ⋅ r + b ⋅ r1
1
2
) (7)
The leading and trailing edges are drawn by applying a deformation function defined by two Bezier curves
consisting of four (4) control points each. Both edges are kept as they have optimized, in a previous step. [6]
The numerical analysis results and comparison of six (6) impellers with blades of varying length, forming
passages with various wrap angles θw is presented below. A characteristic parameter of blade geometry design
and construction is the angle of overlap θ between two consecutive blades (Fig. 2). The overlap angle θ for each
blade can be expressed as function of the wrap angle θw from the equation:

θ = θw − (8)
z
where z denotes the number of the blades. An increase of angle θ results to a longer blade due to the increase of
θw and in contrast, decrease of θ leads to shorter blades (Fig. 4).
Vasilios A. Grapsas, John S. Anagnostopoulos, and Dimitrios E. Papantonis.
The computational domain is a periodically symmetric circular sector of 2π z angle. The impeller
dimensions are regulated according to a commercial centrifugal pump already operating in the Laboratory. The
impeller has, as already said, nine (9) two-dimensional (non-twisted) blades with inlet and exit diameters D1=70
mm, D2=190 mm respectively, exit width b2=9 mm, and inlet and exit angles β1=26 deg, β2=49 deg, respectively.
The blade thickness is constant (5mm) and the nominal speed of rotation is fixed at 3000 rpm.

Blade angle, β
Circumferential

angle, θ
dr
r

Figure 1. Blade angle distribution. Figure 2. Angle of overlap θ of impeller blades.

2.3 Impeller head and power calculations


The energy gained by the fluid through the impeller, H, is computed from the total energy of the fluid at the
impeller inlet and exit periphery (Fig. 3):
⎛ p − p 1 c 22 − c12 ⎞

1
H = H 2 − H1 = ⋅ ⎜⎜ 2 + ⎟ dq (9)
Qu ⎝ ρg 2 g ⎟⎠
where c is the absolute velocity of the fluid, Qu the flow rate through the impeller and g the gravity acceleration,
while the subscripts 1 and 2 denote impeller inlet and exit conditions, respectively (Fig. 3). The head given by
the impeller, Hu, can be calculated from the torque Mu developed on the blades:
r2

∫ [(
G G G G
N u = ρ ⋅ g ⋅ Qu ⋅ H u = ω ⋅ M u = ω ⋅ r × n ) ⋅ p + (r × τ w ) ⋅ cot β ]⋅ b ⋅ dr (10)
r1
G G G
where τw is the turbulent shear stress, n is the normal vector to the blade surface at the location r , and β and
b the local blade angle and width, respectively.

w2
α
u2 2 c
2

w1
α1
u1 c1
r2 r1
ω

Figure 3. Sketch of a centrifugal pump impeller.


Hence the impeller head Hu can be computed as :
ω ⋅Mu (11)
Hu =
ρ ⋅ g ⋅ Qu
and the hydraulic efficiency of the impeller is finally defined as:
H
ηu = (12)
Hu
Vasilios A. Grapsas, John S. Anagnostopoulos, and Dimitrios E. Papantonis.

2.4 Numerical optimization


In order to find the wrap angle θw that maximizes the target value (the hydraulic efficiency), an optimization
procedure is needed. The minimization/maximization of the objective function can be reached using different
techniques. In this study the maximization of the pump efficiency, is realized through an optimization algorithm
based on evolutionary strategies. The Evolutionary Algorithms SYstem (EASY) optimization software [4] used in
this work is a recently developed and brought to market by the Laboratory of Thermal Turbomachinery, NTUA.
It makes use of stochastic optimization methods and has been successfully applied in various engineering
applications for inverse design. The maximization of the impeller hydraulic efficiency ηu, which is calculated
after convergence of the flow solver via equations (9)-(12), will lead to the optimum design.

3 RESULTS

3.1 General
In a preliminary examination of the effect of the wrap angle θw on the hydraulic efficiency of the impeller,
six (6) different geometries are selected, as shown in figure 4. The variation of the wrap angle θw is achieved, by
uniformly changing the “RBM” factor (Table 1, eq. (5)), in order to form different blades, from an almost radial
to a more rounded one.

Impeller RBM θw θ ηu
No
1 1.6 54.1 14.1 0.857
2 1.4 58.5 18.5 0.861
3 1.2 68.6 28.6 0.858
4 1.0 77.8 37.8 0.894
5 0.8 98.1 58.1 0.879
6 0.6 122.6 82.6 0.742

Figure 4. Comparison of blades shape. Table 1 : Impeller geometric data and efficiency.

Performing the corresponding solutions, an optimum in terms of efficiency can be revealed (Fig. 5 and Table
1) for Best Efficiency Point (BEP) flow rate (Q=62,5m3/h). However, in order to predict the performance of the
impeller within a flow rate range around BEP, where it might operate, the construction of the complete ηu-Q
characteristic curve for each case is needed. Six such resulting curves corresponding to the blades of figure 4 are
plotted in Figure 6. It is clearly shown (Fig. 6) that the impeller No 4 (with wrap angle θw=77,8o) displays the
best performance around BEP.
The computed curves Hu–Q and pmin-Q, for the various cases of blade lengths are drawn in figures 7 and 8.
The predicted results in these cases are presented over the entire flow range. The minimum pressure pmin is
calculated with regard to a reference pressure around the blade leading edge. It is remarkable that the computed
H-Q curves for all the different blades tested do not present substantial differences over the entire flow range
and the minimum pressure distribution presents a stable behavior below the BEP. This is because the minimum
pressure is strongly related with the leading edge geometry and the head is related with the β2 angle, which are
both fixed for all cases. However, for flow rates higher than nominal and for θw≥98o a great reduction in
minimum pressure is observed (Fig. 8), because the passage among the blades becomes very thin and the fluid’s
velocity through it increases. For this reason, the efficiency is also much reduced in this case (Fig. 6).
Vasilios A. Grapsas, John S. Anagnostopoulos, and Dimitrios E. Papantonis.

0.92

0.88

0.84
ηu

0.8

0.76

0.72
40 60 80 100 120 140

θw (deg)

Figure 5. Correlation between θw and ηu. Figure 6. ηu-Q computed curve.

Figure 7. Hu-Q computed characteristic Figure 8. pmin-Q computed characteristic curve.

In figure 9a to 9f the flow field at the BEP for the six blades tested is illustrated. In the impeller number 1
(Fig. 9a) the blade length is small and so the small wrap angle does not guide the fluid flow properly and creates
important recirculation. For the impellers number 4 to 6, where the passages become long and the angle of
divergence small (Fig. 9d to 9f), the blade length is larger and so the friction losses, but the recirculation
disappears.
The region of the minimum pressure at Figures (a)-(d) is always located at the blade’s suction side where the
flow accelerates. On the contrary, in figures (e) and (f) the minimum pressure region appears at the pressure side
of the leading edge, because of the very thin passage, which as mentioned above accelerates the fluid flow and
reduces the pressure.
Vasilios A. Grapsas, John S. Anagnostopoulos, and Dimitrios E. Papantonis.

Figure 9. Flow Streamlines and Pressure contours for impellers of various θw angles: a) θw=54,1 , b) θw=58,5 , c)
θw=68,6 , d) θw=77,8 , e) θw=98,1 , f) θw=122,6.

The impeller No 4, with θw=77,8o, which from the above results exhibits the best behavior, has an angle of
overlap θ=38o within the range of the most efficient blade geometry (30-45o) found in an earlier experimental
study for a similar impeller. [13]

3.2 Optimization Procedure

3.2.1 Blade Length


The objective of the optimization is to maximize the impeller’s efficiency, using as design variable the wrap
angle, which expresses the blade length. An additional parameter is the desired operation range of the impeller.
The target here is to maximize the efficiency only at the BEP. The resulting optimal geometry from the above
study is sketched in Figure 10, in comparison with the tested blade No 4. The shape of the two blades is different
and the impeller blade length seems to be an important parameter in order to determine the optimal design in
terms of efficiency. The improved characteristics for the optimal design (θw=72,62o) with achieved efficiency
ηu=0.92 are presented in Figure 11, in comparison with the efficiency obtained with the blade No 4. As
expected, the calculated characteristic curve ηu is in compliance with the parametric study conducted previously
and clearly higher around BEP (approximately 2,5%).

Figure 10. Comparison of blade’s shape. Figure 11. Comparison between obtained efficiency
with optimum geometry and blade No4.
Vasilios A. Grapsas, John S. Anagnostopoulos, and Dimitrios E. Papantonis.
3.2.2 Number of blades
The number of impeller blades has been initially selected equal to z=9. The effect of using more or fewer
blades can be explored easily with the performance calculation computer program. More blades guide the flow
better, but the friction losses are increased and the blockage produced by the blade thickness at the inlet becomes
stronger. At first, a preliminary study for different blade numbers (z=6, 7, 8, 9, 10 and 11) is presented in
Figures 12 and 13, obtained with the numerical code and through the variation of the wrap angle for the entire
flow range. In the results of Fig. 12, maximum efficiency exhibits the impeller with 9 or more blades, whereas
the BEP is for all impellers at the same, nominal flow rate. In figure 13 the ηu-θw curves are obtained for this
nominal flow rate. As can be observed, the impeller with z=9 blades tends to present the most efficient behavior.

Figure 12. ηu-Q computed characteristic curve. Figure 13. ηu-θw computed curve.
By applying the optimization algorithm for the impellers with blade number z=6, 7, 8, 10 and 11, the
optimum blade shape, which maximizes the hydraulic efficiency for each blade number is revealed (Fig. 14).
The angle of overlap as well as the wrap angle tends to vary within a restricted “optimal” region (27o<θ<37o)
(Table 2). From the optimization procedure it is extracted that the impeller with 9 blades presents the most
efficient behavior (Fig. 15). As can be observed from both Figures 14 and 16, an increase of the number of
blades z corresponds to a gradual decrease of the wrap angle θw and thus to the blade’s length. This can be
explained from the fact that an increase of z leads to thinner passages among the blades and higher flow
velocities. In order to decrease the blockage effect and the hydraulic losses, the blade’s length as well as the
wrap angle must be reduced.

Optimu Optimum
ηu m θw z θ
0.894 68.96 11 36.23
0.907 71.09 10 35.09
0.919 72.63 9 32.63
0.889 74.29 8 29.29
0.881 82.64 7 31.21
0.872 87.96 6 27.96

Figure 14. Optimum blades for various Table 2 : Optimum angles θw, θ, and best efficiency for
blade numbers. the six examined impellers.
Vasilios A. Grapsas, John S. Anagnostopoulos, and Dimitrios E. Papantonis.
0.95 90

0.925 85

0.9 80
ηu

θw
0.875 75

0.85 70

0.825 65

0.8 60
5 6 7 8 9 10 11 12 5 6 7 8 9 10 11 12

z z
Figure 15. Maximum ηu-z correlation Figure 16. θw-z for maximum impeller efficiency

4 CONCLUSIONS
A numerical method for the blade’s hydrodynamic design of a radial flow impeller is developed and
evaluated, combining an automated fluid flow solver, a shape parameterization technique and an optimization
software. The method is applied here to improve the efficiency of a centrifugal pump impeller by optimizing the
blade’s length and the number of blades for given operating conditions and other geometric data. The optimum
blade shape is obtained quite fast and exhibits a considerably improved performance over a wide operating
range around BEP. It was found that the optimum blade length decreases as the number of blades increases. The
exact regulation of the design parameters, namely the blade wrap angle and the number of blades, was found to
improve the hydraulic efficiency of the impeller by several percentage units above a standard design
performance. The developed methodology can be applied to optimize all the other geometric parameters of a
centrifugal pump impeller.

ACKNOWLEDGEMENTS
The project is co-funded by the European Social Fund (75%) and National Resources (25%)-Operational
Program for Educational and Vocational Training II (EPEAEK II) and particularly the Program Pythagoras.

REFERENCES
[1] Japikse, D. (1994), “Advanced turbomachinery design methods”, Computational Fluid Dynamics for
Turbomachinery Design (Concepts ETI, Inc., Wilder, Vermont).
[2] Ferrando, L., Kueny, J. and Avellan F. (2004), “Surface Parameterization of a Francis Runner Turbine for
Optimum Design”, Proceedings of the 22nd IAHR Symposium on Hydraulic Machinery and Systems,
Stockholm, Sweden, Jun 29-July2.
[3] Adrizzon, G. and Pavesi, G. (1995), “Theoretical evaluation on the effects of the impeller entrance
geometry and the incident angle on cavitation inception in centrifugal pumps”, Proc. Instn. Mech. Engirs.
Vol. 209, pp. 29-38.
[4] Giannakoglou, K.C. (2002), “Design of optimal aerodynamic shapes using stochastic optimization methods
and computational intelligence”. Progress in Aerospace Sciences, Vol 38, pp. 43-76.
[5] Benra, F.K. (2001), “Economic development of efficient centrifugal pump impellers by numerical
methods”, World Pumps, May 2001, pp. 48-53.
[6] Grapsas, V., Anagnostopoulos, J. and Papantonis D. (2005), “Hydrodynamic Design of Radial Flow Pump
Impeller by Surface Parameterization”, Proceedings 1st IC-EpsMsO, Athens, Greece, 6-9 July.
[7] Grapsas, V. and Anagnostopoulos, J. (2004), “Numerical optimization of the hydrodynamic shape of fluid
flow systems”, Proceedings 1st IC-SCCE, Athens, Greece, 8-10 September.
[8] Anagnostopoulos, J. (2003), “Discretization of transport equations on 2D Cartesian unstructured grids
using data from remote cells for the convection terms”, Intl. J. for Numerical Methods in Fluids, 42, pp.
297-321.
[9] Ye, T., Mittal, R., Udaykumar, H.S. and Shyy, W. (1999), “An accurate Cartesian grid method for viscous
incompressible flows with complex immersed boundaries”, J. of Computational Physics, 156, pp. 209-240.
[10] Fadlun, E.A., Verzicco, R., Orlandi, P. and Mohd-Yusof, J. (2000), “Combined immersed-boundary finite-
difference methods for 3D complex flow simulations”, J. Computational Physics, 161, pp. 35-60.
[11] Anagnostopoulos, J. (2005), “A numerical Simulation Methodology for Hydraulic Turbomachines”, 5th
GRACM International Congress on Computational Mechanics, Limassol, Cyprus, 29 June-1 July.
[12] Stepanoff, A.J. (1993), “Centrifugal and Axial Flow Pumps, Theory, Design and Application”, Krieger
Publishing Company Malabar, Florida.
Vasilios A. Grapsas, John S. Anagnostopoulos, and Dimitrios E. Papantonis.
[13] Lazarkiewicz, S. and Troskolanski, A. (1965), “Impeller Pumps”, Pergamon Press, Warsaw.

Вам также может понравиться