You are on page 1of 12

International Journal of Fluid Machinery and Systems DOI: http://dx.doi.org/10.5293/IJFMS.2016.9.4.

313
Vol. 9, No. 4, October-December 2016 ISSN (Online): 1882-9554

Original Paper

Centrifugal Impeller Blade Shape Optimization Through Numerical


Modeling

Sayed Ahmed Imran Bellary1 and Abdus Samad2


1
Department of Mechanical Engineering, ADCET Ashta, Shivaji University Kolhapur
Sangli-416 301, Maharashtra, India, ibellary@gmail.com
2
Indian Institute of Technology Madras Chennai-36, India, samad@iitm.ac.in

Abstract

Surrogate model based shape optimization methodology to enhance performance of a centrifugal pump has been
implemented in this work. Design variables, such as blade number and blade angles defining the pump impeller blade
shape were selected and a three-level full factorial design approach was used for efficiency enhancement. A three-
dimensional simulation using Reynolds-averaged Navier Stokes (RANS) equations for the performance analysis was
carried out after designing the geometries of the impellers at the design points. Standard k-ε turbulence model was used
for steady incompressible flow simulations. The optimized impeller incurred lower losses by shifting the trailing edge
towards the impeller pressure side. It is observed that the surrogates are problem dependent and most accurate surrogate
does not deliver the best design always.

Keywords: Centrifugal impeller, Optimal design, Blade number, Blade angles, Multiple Surrogate models, Hydraulic
efficiency.

1. Introduction
Centrifugal pump is used for fluid transportation in almost all industries and belongs to a class of turbomachinery. The centrifugal
action of impeller blade converts mechanical energy into hydraulic energy. Performance characteristics of pump depend on the impeller
geometry. For better performance of pump, design parameters like impeller blade width, exit blade angle, blade height, number of
blades, impeller diameter and curvature of blade plays key role.
The performance analysis of pump or impeller was studied with the help of computational fluid dynamics (CFD) simulations [1-2].
To obtain thorough performance prediction of a specific design on fluid behavior in the machine, numerical simulations is an alternate
choice and CFD analysis predicts acceptable solutions within limited time [2-4].
Several authors [5-8] have reported the effect of number of blades on the centrifugal pump. Feng et al. [7] performed numerical
investigation for different shaped vane diffuser pumps and reported that increase in the number of the impeller blades leads to an
increase of pressure fluctuation on the impeller pressure side and on the whole diffuser vane surface. Rababa [8] performed both
experiments and simulations on a centrifugal pump with different blade numbers and shape to investigate the flow field and found that
increase in blade number augmented the head generation.
Sanda and Daniela [9] performed a numerical study on the radial impeller as a function of inlet blade angle and achieved an
efficiency increment by 1-2%. Luo et al. [10] performed experimental test and numerical simulation to understand the effect of inlet
geometry on centrifugal pump and reported that hydraulic performance of pump was increased. There are experimental or numerical
results for different discharge angles for different centrifugal impellers [11-12] and they have shown an improvement in efficiency.
Viscosity which can be defined as flow resistance reduces performance [13].
Surrogate approximation models are widely used in the design and optimization problems and it has been observed that the surrogate
models perform differently for different applications [14]. Goel et al. [14] developed an averaged surrogate (PBA) models using
response surface approximation (RSA), radial basis neural network (RBNN), and Kriging (KRG) models. Samad [15] implemented the
method and concluded that the weighted average surrogate model was much reliable prediction method than an individual surrogate.
Recently Derakhshan et al. [16] used optimization method for the pump impeller based on the artificial neural network during
optimization phase for application of artificial bee colony algorithms. Geometrical parameters like hub diameter, suction diameter,
impeller diameter, impeller width and inlet and outlet blade angles were chosen as the optimization variables for the enhancement of
efficiency. They reported 3.59% augmentation in efficiency for the optimized pump.
In the present problem; impeller blade number, inlet and exit angles were design variables and hydraulic efficiency was an objective
function. Different surrogate models combined with a three-dimensional Navier-Stokes solver was used and flow properties were
evaluated. Detailed description of the methodology and flow analysis has been presented.

Received March 5 2015; revised December 09 2015; accepted for publication August 5 2016: Review conducted by Young Seok Choi.
(Paper number O15007K)
Corresponding author: Abdus Samad, Associate Professor, samad@iitm.ac.in

313
2. Numerical Formulation
2.1 CFD Modeling

The geometry and geometric parameters of impeller are shown in Fig 1 Table 1, respectively. For investigating the influence of
blade number and blade angles on the performance of centrifugal impeller three different blade numbers 6, 7 and 8, inlet angles
10o, 15o and 20o and three outlet blade angles 30o, 40o and 50o are selected and optimization has been performed. Table 2
represents the three different fluids with density and dynamic viscosity used for analysis. All the simulations have been
performed at 20o C fluid temperature.

Table 1 Features of impeller

Parameter Dimension
Shaft diameter, Ds 40 mm
Eye diameter, Do 182 mm
Hub diameter, Dh 55 mm
Inlet diameter, D1 160 mm
Inlet blade width, b1 54 mm
Outlet blade width, b2 30 mm
Inlet blade angle, β1 10o -20o
Outlet blade angle, β2 30o-50o
Blade number, z 6-8
Blade thickness, y 5 mm
Outlet diameter, D2 365

Geometrical modeling and meshing of the flow domain was carried out by using Ansys-BladeGen and TurboGrid module. Ansys-
CFX 13.0 [17] was used for the flow simulations and the problem setting is given in Table 3. The mesh in the flow passage is with O-
grid block and at the inlet and outlet it is with H-grid block. These grid blocks, help to account the wall shear stress and boundary layer
losses accurately [18]. The mesh has been shown in Fig. 1(c).

Table 2 Viscosity and density of the fluids

Fluid Viscosity [N-s/m2] Density [kg/m3]


Water 1.002E 997
Crude-oil 5.000E 835
Gasoline 5.000E 720

Table 3 Meshing and boundary conditions


Flow domain Single impeller
Interface Periodic
Mesh/Nature Structural/Hexahedral
Nodes/Elements 634,161/568,620
Fluid used Water, Crude oil and
Gasoline
Turbulence model k-ε
Inlet/Outlet Pressure/Mass flow rate
Residual convergence 1x10-5
Residual type RMS
Iteration steps 2000
Mass imbalance % 0.0001

314
X
W1
cm1= c1
β1 α1
Y Wu1 Z
u1
A

W2
c2
cm2
β2 α2
B Wu2 D cu2 C
u2
(a) Inlet and exit velocity triangle

Hub
Inlet

Blade

(b)
I ll
Periodic

Blade

Shroud Inlet

Hub

Outlet
(c) Flow domain and mesh

(d) Leading edge (e) Trailing edge

Fig. 1 Inlet and outlet velocity


triangle, geometrical model and
flow domain

315
To account for Reynolds number (Re>105) standard k-ε turbulence model was employed. The k-ε turbulence model is a class of
RANS models to which two partial differential equations are solved, one for turbulence kinetic energy, k, and another for turbulence
kinetic energy dissipation rate, ε, to get length and time scale information needed to form local eddy viscosities [19]. The simulations
were carried out on 3.4 GHz core i7-3370 processor with 8 GB ram. Time per iteration was within 29 seconds.
The blades were swept backwards (β2 < 90o) and used for pumps and blowers [20]. The blade exit angle exerts a decisive influence
on the total head and efficiency of pump. As per Euler's equation head generated by impeller blade is given by:

cu 2u 2 − cu1u1
H= (1)
g

Normal entry at impeller inlet implies cu1=0 and from velocity triangle shown in Fig. 1(a) cu2 can be given as

Q cot b 2
cu 2 = u 2 − (2)
πDb2

From Fig. 1(a) it is clear that as β2 increases, the absolute exit velocity c2 also increases. Increase in c2 causes cu2 to increase which
results in increase in head.
Hydraulic efficiency is given by,

H a H th − ∆H
h= = (3)
H th H th

2.2 Optimization Methodology

The objective of the problem was to enhance hydraulic efficiency (η) (eq. 11) and the design variables were blade number (z), inlet
blade angle (β1) and exit angle (β2). The authors [3, 7-10, 21] found that these are the most important parameters to increase
performance of an impeller. Ranges of the variables (Table 4) are selected based on preliminary calculations and prior literature search
[3, 21-22, 23-24]. A three-level full factorial design was used to select 27 design points within the variable ranges. These points were
used to get efficiency values through CFD analysis and the values were used to construct the surrogate models.
Optimization methodology deals with the selection of the best alternative in the sense of the given designs. The procedure for
surrogate based optimizations is presented in Fig. 2. In order to evaluate an objective function as a function of design variables, a lot of
numerical simulations are required. For most of thermo-fluids problems, however, a single numerical simulation, such as 3D RANS
analysis takes long time to complete. Surrogate models mitigate this burden by constructing approximation models that imitate the
behavior of the simulation model as closely as possible while being computationally cheaper to evaluate. The surrogates are also
problem dependent which creates difficulty in selecting any surrogate for a particular problem. The surrogates, such as RSA, RBNN,
KRG and PBA were implemented by Samad and Kim [25] which are also used for this problem. A weighted average surrogate model
was constructed and a search algorithm was used to find optimal points from the constructed surrogates using a sequential quadratic
programming.

Problem statement Design of experiment Numerical analysis Formulation of


Objective functions, design Allocation of design Calculation of the values of objective surrogates
space and design variables points functions at each experimental point RSA, RBNN, KRG or PBA

No

Whether Quest for optimal point


Design of experiment Yes optimal design point within Optimal point quest from constructed
Allocation of design points
design space? surrogate using optimization algorithm

Fig. 2 Optimization procedure

316
Table 4 Design space of the variables

Variables Lower limit Upper limit


β1 10 20
Β2 30 50
z 6 8

3. Results and Discussions


3.1 Grid Independency and Validation, Turbulence Models

The flow rate, head and rotation of impeller meet the specifications given in [21]. Reasonably good agreement is seen between the
simulated values for water, as fluid, with that of the analytical solutions as in ref. [21]. The computations were carried out at different
flow rates, i.e at design and off-design conditions. Figure 3 represents the grid independency test conducted for water at designed flow
rate.

46

44
H

42
2 3 4 5 6
Number of nodes (X10-5)

Fig. 3 Head variation with number of nodes at design flow rate (Q=0.1003 m3/s)

The pump head at design flow rate was taken as the parameter to evaluate grids and the influence of mesh size on solution was
calculated by setting convergence criteria of 10-5. Once the number of nodes reach a value of 550,000, the variation of head with the
increase in number of nodes was not significant.
The results of CFD simulations were compared with the analytical results at design point when water was used as working fluid. The
head and efficiency developed was over estimated by 7% and 2% as compared to that of the analytical result. The present result is
similar to that of the others [26-27] and a difference of about 10% result between CFD and analytical is allowed [28]. Simulations were
performed at 1470 rpm for design and off design conditions to study flow pattern.
The turbulence models are still on developing stage and it is required to test the models for different problems, which can produce
more accurate result. Due to flow complexity, the choice of turbulence model is delicate issue to apply in flow simulations, for
turbomachinery systems. Bradshaw [29] has reported that the selection of turbulence model is influenced by three criteria: a) physical
nature of problem, b) quality of attended results and c) computing power. In spite of this, the traditional RANS models, like k-ε or k-ω
are largely used and satisfactory results are achieved. The models k-ε, SST and k-ω were evaluated for impeller flow simulation, at
design point and under similar conditions. Analytical result produced 40 m head at design flow rate, whereas by using k-ε, SST and k-ω
turbulence models the values obtained are 43.11m, 43.95m and 45.15m, respectively. The least error (=7.21%) among the turbulence
models was produced by k-ε and this model was chosen for simulations.

3.2 Optimization Results

Initially parametric study was performed for the impeller and it was found that the impeller with z =7 has slightly better performance
as compare to z =6 and 8 designs, hence all further calculations were performed at z =7 with different β1 and β2 angles. Tables 5 contain
the optimization results for the objective function with different surrogate models. It was found that the results show that the RSA gives
lowest CV error among the other surrogates, and the RBNN performs the worst for the objective function. The low cross validation
(CV) error for RSA is likely to indicate that the objective is modeled well as quadratic polynomials [30]. In PBA, the weight associated
with RSA is the largest. Hence the equation for the PBA model takes a form:

Fpba= 0.388 Frsa+ 0.301 Frbnn+ 0.312 Fkrg (4)

317
Table 5 Optimal designs suggested by various surrogates and corresponding predicted RANS results for objective function

Surrogate model RSA RBNN KRG PBA

Optimal design β1 16.02 16.07 15.38 15.73


variables β2 47.52 43.91 42.00 43.96
Fsur 92.15 92.22 92.17 92.16
FRANS 92.09 92.05 92.06 92.09
FRANS - Fsur -0.06 -0.16 -0.18 -0.08
Fref 90.00 90.00 90.00 90.00
FRANS - Fref 2.09 2.06 2.06 2.09
Increase in efficiency 2.32 2.29 2.29 2.32

Table 5 compares the predictions of the optimal values predicted by different surrogate models against the actual RANS calculation
at the corresponding points. Consistent with the aforementioned error estimate, RBNN performs the worst. RSA and PBA give better
prediction of objective function value. Depending on the characteristics of the design space in each problem, certain surrogate models
can perform more satisfactorily than others. However, it is difficult to tell a priori which surrogate model is more appropriate than the
others. These are precisely the merits of adopting multiple surrogates and finding ways to form a more robust composite surrogate [30].
The Table 5 shows the CFD and surrogate predicted results. The equation (4) was used to find prediction by PBA model. The highest
efficiency enhancement was found by RSA and PBA (=2.32%). It can also be noticed that the error in prediction (=FRANS-Fsur) is lower
in the case of RSA and PBA at the optimal points predicted by the respective surrogates. Figure 4 shows the sensitivity of objective
function with the 10% deviation of variables from optimal point. RSA model is used to evaluate objective function values. This figure
shows that the outlet blade angle is more sensitive to change its value and a small change in value may lead to larger change in
efficiency.

Fig. 4 Sensitivity of the variables

Fig. 5 Surrogate predicted response surface for RSA

318
A response surface generated using the best surrogate RSA. Fig. 5 shows that the optimal values of the variables and the efficiency.
Normalized values (ranged from 0 to 1) of the variables are presented in the figures. Peak value of the surface shows optimal objective
function value.
The optimum values of β1 and β2 are 16.02 and 47.52, respectively. The optimal point having peak hydraulic efficiency (92.15) is
located inside the boundary of design space as shown in Table 4. Here, one point can be noticed that the optimal objective function
values are not hitting the boundary, which helped to accept the result in single optimization iteration. If improper design space was
selected a reiteration was necessary to find the objective function value in the design space.

3.3 Performance Analysis

The typical oil industry handles different viscosity fluids. In the present investigation, three fluids are considered for the performance
evaluations (Table 2). The reference and optimized geometry for water was used for the evaluation of pump performance for crude oil
and gasoline. The performances were compared for all the fluids given in Table 2.

Fig. 6 Characteristic curves for different


fluids at design and optimum point

Performance curves for the fluids at design and off-design conditions are shown in Fig. 6. The hydraulic losses and theoretical
head at design point differ to small extent from off design conditions. A measure of hydraulic efficiency is directly linked with
head generated. As the head measurement is high for large exit blade angle, the corresponding hydraulic efficiency is more for the
liquid with large discharge angle [24]. The optimized β2 has large exit blade angle value compare to designed blade. This results in
an increase in head generation. It is due to increase in absolute flow velocity at exit (c2) and corresponding peripheral velocity (cu2)
i.e. dynamic part of head increases more rapidly shown in Figure 1 (a). This is evident from Eq. (1) and (2).
Figures 7 demonstrate the contours of static pressure for design point (Q=0.1003 m3/s) at 50% span. It is clear from these figures that
static pressure difference between impeller inlet and outlet increases with large exit blade angle. It is due to the fact that fluid flow
velocity at impeller outlet decreases with increase in discharge angle. The diameter at the outlet is high; hence absolute velocity as well

319
REF blade
Water Crude oil Gasoline

PS
SS

OPT blade
Rotational
direction

Fig. 7 Static pressure contours at 50% span for


0.1 m3/s for RSA

REF blade
Water Crude oil Gasoline

OPT blade

Fig. 8 Meridional view of velocity


contours (Q = 0.1 m3/s) for RSA

320
as total pressure at the outlet is augmented. In addition, the increase in discharge angle increases the total pressure at the outlet. This
enables the fluid velocity at the outlet to be decreased.
The pressure contours represents a smooth flow between the blades and its value increases continuously towards the exit of the
computational domain. It is to be noted that the lowest static pressure was observed at the impeller inlet on suction side which is a
location in centrifugal impeller where usually cavitation appears. Here it is above saturation pressure of liquid and the occurrence of
cavitation is ruled out [27-28].
The highest static pressure occurs at the impeller exit and the kinetic energy of the flow reaches to its maximum extent. Key factor is
the pressure which increases continuously as the mechanical energy given in the form of impeller rotation is converted into the pressure
energy. Minimum pressure exists at the suction side and near the leading edge of the blade. The difference in pressure between outlet
and inlet increases with increase in blade angle. Liquids at large exit blade angle (optimized blades) have more difference in pressure
between inlet and outlet. The peripheral velocity (cu2) at the impeller outlet elevates for large discharge angle as depicted from Fig.1 (a).
Reasonable good agreement is seen between the existing results and those of [23, 26-27]. Head being influenced by peripheral velocity
at out let increases and in turn hydraulic efficiency as well.
The hydraulic efficiency for the handling crude oil and gasoline is lower than that for the handling water. The decrease in efficiency,
while pumping the crude oil and gasoline is due to the viscosity which results in disc friction losses over the outsides of the impeller
shroud and hub. These results are in good match with the analytical results of Srinivasan [20] and Gulich [28]. In addition to this, the
head and discharge of the pump for which it is designed are at maximum efficiency point where hydraulic losses, profile losses and
secondary losses are minimum. On the contrary at all other point of operation shock losses and secondary losses increases resulting into
great hydraulic losses and decrease in hydraulic efficiency as well.
Figure 8 shows the meridional view for distributions of velocity contours for various fluids at reference point and optimum point for
RSA/PBA. The optimized exit blade angle has large exit blade angle value compare to designed blade. This results in an increase in
head generation. It is due to increase in absolute flow velocity at exit (c2) and corresponding peripheral velocity (cu2), i.e., dynamic part
of head increases more rapidly. This is evident from equations (1) and (2). The highest velocity occurs at the impeller outlet and the
kinetic energy of flow reaches to its maximum extent. The difference in absolute velocity from outlet to inlet increases with increase in
blade angle. Liquids with optimized exit blade angle have more difference in velocity between inlet and outlet and greater velocity at
the outlet. It is due to increase in peripheral velocity (cu2) at the impeller outlet for large discharge angle as reflected from Figure 1(a).
This agrees the analytical results; as fluid flows from inlet to outlet of the impeller, the absolute velocity and pressure increases and for
large exit blade angle (optimized angle) its magnitude is high.

REF blade LE
Water Crude oil Gasoline

OPT blade

REF blade TE

OPT blade

Fig. 9 Total pressure (kPa) contours at LE on SS and at TE


on PS for reference and optimized blade

321
Figure 9 shows the distributions of total pressure on suction surface at leading edge of designed and optimum blade. Pressure
changes along the leading edge on the suction surface of the blade. And, in the case of the efficiency optimized blade, it is noted
that near tip of leading edge on suction surface the pressure shows an increased value. Therefore, it is found from this figure that
enhancement of efficiency is achieved by modifying the flow angle near the tip of the blade. This is consistent with the results of
Gulich [28], who suggested that the optimization of inlet blade angle has remarkable influence on efficiency. Figure 9
demonstrates the total pressure contour on pressure side at trailing edge of designed and optimized blade. It is observed that there
is an increase in the total pressure along trailing edge of the blade. For optimized blade increase in total pressure along trailing
edge is large compare to designed blade. This in turn increases the head and consequently increases the hydraulic efficiency.
Figure 10 depicts the magnitude and nature of peripheral velocity from inlet to outlet at reference and optimized blade for RSA.
It is clear that the optimized blade for all the fluids have higher peripheral velocity component (cu2) at the outlet as compared to
reference blades. It is because the absolute velocity and total pressure at the outlet is higher for large exit blade angle.

REF blade
Water Crude oil Gasoline

OPT blade

Fig. 10 Peripheral velocity distributions. (Q =


0.1 m3/s) for RSA

2. Conclusion
The shape optimization of the blades for centrifugal pump impeller was performed for single objective function with three-
dimensional Navier-Stokes analysis. The conclusions are:

• The optimized blade shape produced higher hydraulic efficiency, and this is due to the reduction in flow loss and the disc
friction losses over the impeller shroud and hub. The higher viscosity fluids have higher friction losses which hinders the
impeller performance.
• At optimized point; hydraulic losses, profile losses and shock losses were found to be lease resulting into an increase in
absolute flow velocity at the exit. This has facilitated to increase in dynamic part of head and consequently an augmentation
in head has been observed.
• Increase in outlet blade angle by shape optimization has enabled to an increase in outlet cross section size, hence a reduction
in fluid pressure drop is observed. This helps in a relatively high head generation and also in efficiency enhancement.
• RBNN gives the highest cross-validation error and performs the least for the objective function. RSA, with the less cross-
validation error, shows the best reliability of the surrogate. The weighted surrogate PBA has produced satisfactory result
which can be used reliably in the case of pump design modification and protects against choosing a poor surrogate if only one
is chosen.

322
Acknowledgments
The authors would like to acknowledge Indian Institute of Technology Madras for the NFSC grant (Grant code: OEC/10-
11/529/NFSC/ABDU) to conduct this research.

Nomenclature
2
Ac Cross-section area of micro-channel [m ] T Local mean temperature [K]
Lx Length of heat sink [m] Ui Mean velocity components (i=1, 2, 3)
Rth Thermal resistance [ oC/W] θ, φ Design variables, Wc/Hc and Ww/Hc
Re Reynolds number (=UbDh/ν) ρ Fluid Density
b Breadth, mm a Actual
CFD Computational fluid dynamics design Design flow
CV Cross validation h Hub
c Absolute fluid flow velocity, m/s krg Kriging
D Diameter, mm m Meridional component
g Acceleration due to gravity, m/s2 o Eye
H Head, m pba PRESS based averaging
∆H Hydraulic losses, m RANS Reynolds-averaged Navier Stokes
KRG Kriging RBNN Radial basis neural network
k Turbulence kinetic energy, J RSA Response surface approximation
LE Leading edge s Shaft
OPT Optimized sur Surrogate
P Power, W th Theoretical
PBA PRESS based averaging u Peripheral component
PS Pressure side w Relative fluid velocity component, m/s
p Pressure, N/m2
Q Volume flow rate, m3/s Greek Symbols
RANS Reynolds-averaged Navier Stokes α Flow angle, o
RBNN Radial basis neural network β Blade angle, o
REF Reference ε Rate of energy dissipation, J/s
RSA Response surface approximation η Hydraulic efficiency, %
SS Suction side μ Dynamic viscosity, N-s/m2
SST Shear stress transport ρ Density of fluid, kg/m3
TE Trailing edge ω Angular velocity, rad/s
u Peripheral velocity, m/s
W Relative fluid velocity, m/s

Subscript
1 Inlet
2 Outlet

References
[1] Zhou, W., Zhao, Z., Lee, T. S., and Winoto, S. H., 2003, “Investigation of Flow through Centrifugal Pump Impellers using
Computational Fluid Dynamics,” International Journal of Rotating Machinery, Vol. 9, No. 1, pp. 49-61.
[2] Zhang, Y., Zhou, X., Ji, Z., and Jiang, C., “Numerical Design and Performance Prediction of Low Specific Speed Centrifugal
Pump Impeller,” International Journal of Fluid Machinery and Systems, Vol. 4, No. 1, pp. 133-139.
[3] Kamimoto, G., and Matsuoka, Y., 1956, “On the Flow in the Impeller of Centrifugal Type Hydraulic Machinery (The 2nd
Report),” Transactions of the JSME, Series 3, Vol. 22, No. 113, pp. 55-59.
[4] Samad, A., and Nizamuddin, M., 2013, “Flow Analyses inside Jet Pump used for Oil Wells,” International Journal of Fluid
Machinery and Systems, Vol. 6, No. 1, pp. 1-10.
[5] Murakami, M., Kikuyama, K., and Asakura, E., 1980, “Velocity and Pressure Distributions in the Impeller Passages of
Centrifugal Pump,” ASME J. of Fluids Eng., Vol. 102, No. 4, pp. 420-426.
[6] Houlin, L., Yong, W., Shouqi, Y., Minggao, T., and Kai, W., 2010, “Effects of Blade Number on Characteristics of
Centrifugal Pumps,” Chinese Journalof Mechanical Engineering, Vol. 23, pp. 1-6.
[7] Feng, J., Benra, F. K., and Dohmen, H. J., 2007, “Numerical Investigation on Pressure Fluctuations for Different
Configurations of Vaned Diffuser Pumps,” International Journal of Rotating Machinery, Volume 2007, Article ID
34752,10pages.doi:10.1155/2007/34752.

323
[8] Rababa, K. S., 2011, “The Effect of Blades Number and Shape on the Operating Characteristics of Groundwater Centrifugal
Pumps,” European Journal of Scientific Research, ISSN 1450-216X, Vol.52, No. 2, pp. 243-251.
[9] Sanda, B., and Daniela, C. V., 2012, The Influence of the Inlet Angle Over the Radial Impeller Geometry Design Approach
with Ansys, Journal of Engineering Studies and Research, Vol. 18, No. 4, pp. 32-39.
[10] Luo, X., Zhang, Y., Peng, L., Xu, H., and Yu, W., 2008, Impeller Inlet Geometry Effect on Performance Improvement for
Centrifugal Pumps, Journal of Mechanical Science and Technology, Vol. 22, pp. 1971-1976.
[11] Liu, C. H., Vafidis, C., and Whitelaw, J. H., 1994, “Flow Characteristics of a Centrifugal Pump,” ASME Journal of Fluids
Eng., Vol. 116, No. 2, pp. 303-309.
[12] Varley, F. A., 1961, “Effects of Impeller Design and Surface Roughness on the Performance of Centrifugal Pumps,” Proc.
Instn. Mech. Engrs, Vol. 175, No. 21, pp. 955-969.
[13] Li, W. G., 2002, “The Influence of Number of Blades on the Performance of Centrifugal Oil Pumps,” World Pumps, Vol. 427,
pp. 32-35.
[14] Goel, T., Haftka, R. T., Shyy, W. and Queipo, N., 2007, “Ensemble of Surrogates,” Structural and Multidisciplinary
Optimization,” Vol. 33, No. 3, pp. 199-216.
[15] Samad, A., 2008, “Numerical Optimization of Turbomachinery Blade Using Surrogate Models,” PhD Thesis, School of
Mechanical Engineering, Inha University, Republic of Korea.
[16] Derakhshan, S., Pourmahdavi, M., Abdolahnejad, E., Reihani, A. and Ojaghi, A., 2013, “Numerical Shape Optimization of a
Centrifugal Pump Impeller Using Artificial Bee Colony Algorithm,” International Journal of Computers & Fluids, Vol. 81, pp.
145-151.
[17] Ansys-CFX 13.0, 2010, Ansys Inc.
[18] Bellary, S. A. I., and Samad, A., 2015, Numerical Analysis of Centrifugal Impeller for Different Viscous Liquids,
International Journal of Fluid Machinery and Systems, Vol. 8, No. 1, DOI: http://dx.doi.org/10.5293/IJFMS.2014.12.26.00033.
[19] Wilcox, D. C., 1994, Turbulence Modeling for CFD, DCW Industries, Inc., 2nd Edition, La Canada, California.
[20] Srinivasan, K. M., 2008, Rotodynamic Pumps, New Age International (P) Ltd., New Delhi.
[21] Lazarkiewicz, S., and Troskolanski, A. T., 1965, Impeller Pumps, 1st Edition., Pergamon Press Ltd., Oxford.
[22] Stepanoff, A. J., 1964, Centrifugal and Axial Flow Pumps, 2nd Edition., John Wiley and Sons inc., New York.
[23] Aoki, K., Ohta, H., and Nakayama, Y., 1987, “Study on Centrifugal Pump for High Viscosity Liquids (The 1st Report, Effect
of Impeller Output Angle and Number of Blades on the Pump Performance of Closed Type Impellers),” Proceedings of the
School of Engineering, Tokai University, Vol. 27, No. 2, pp. 151-158.
[24] Fard, M. H. S., and Boyaghchi, F. A., 2007, “Studies on the Influence of Various Blade Outlet Angles in a Centrifugal Pump
When Handling Viscous Fluids,” American Journal of Applied Sciences, Vol. 4, No. 9, pp. 718-724.
[25] Samad, A., and Kim, K. Y., 2009, “Application of Surrogate Modeling to Design of a Compressor Blade to Optimize
Stacking and Thickness,” International Journal of Fluid Machinery and Systems, Vol. 2, No. 1, pp. 1-12.
[26] Gupta, M., Kumar, S., and Kumar, A., 2011, “Numerical Study of Pressure and Velocity Distribution Analysis of Centrifugal
Pump,” International Journal of Thermal Technologies, Vol.1, No.1, pp. 117-121.
[27] Bellary, S. A. I., Husain, A., and Samad, A., 2014, “Effectiveness of Meta-models for Multi-objective Optimization of
Centrifugal Impeller,” Journal of Mechanical Science and Technology, Vol. 28, No. 12, pp. 4947-4957.
[28] Gulich, J. F., 2010, Centrifugal Pumps, 2nd Edition., Springer Publications, Berlin.
[29] Bradshaw, P., 1996, “Turbulence Modeling with Application to Turbomachinery,” Progress in Aerospace Sciences, Vol. 32,
No. 6, pp. 575-624.
[30] Samad, A., Kim, K. Y., Goel, T., Haftka, R. T., and Shyy, W., 2008, “Multiple Surrogate Modeling for Axial Compressor
Blade Shape Optimization,” Journal of Propulsion and Power, Vol. 24, No. 2, pp. 302-310.

324

You might also like