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

Journal of Mechanical Science and Technology 29 (5) (2015) 2025~2034

www.springerlink.com/content/1738-494x

DOI 10.1007/s12206-015-0423-4

Influence of stator vane number on performance of the axial-flow pump


Can Kang1,*, Xiaojie Yu1, Weifeng Gong2, Changjiang Li2 and Qifeng Huang3
1

School of Energy and Power Engineering, Jiangsu University, Zhenjiang, 212013, China
2
Shanghai Marine Equipment Research Institute, Shanghai, 200031, China
3
Shanghai Kaiquan Pump Group Co., Ltd, Shanghai, 201804, China

(Manuscript Received September 17, 2014; Revised January 21, 2015; Accepted January 29, 2015)
----------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------

Abstract
The interplay of impeller and stator is investigated for an axial-flow pump. Three stator vane numbers of 5, 7 and 9 are devised to
match the same impeller. The renormalization group k- turbulence model is used to simulate three-dimensional flows in three pumps
with different vane numbers. Axial velocity distributions at impeller outlet and stator outlet are comparatively analyzed. Experiments
assess operation performance of the three pumps. Vibration parameters and static pressure fluctuations are measured as well. It is indicated that the influence of vane number on both pump head and pump efficiency is insignificant. Large stator vane number contributes to
the improvement of the uniformity of axial velocity distribution at impeller outlet. At stator outlet, similar tendency is revealed. Severe
vibration occurs near the outlet bend of the pump, as is particularly remarkable at vane number of 9. For the three cases, blade passing
frequency and its harmonics are predominant in the frequency spectra of pressure fluctuations. As flow rate increases from 0.8Q to 1.0Q,
high-frequency pressure fluctuations are suppressed considerably at vane number of 9, while the other two cases also manifest a decline
in overall pressure fluctuation amplitude. The 7-vane case is the most preferable one among the three cases in terms of both pump performance and pressure fluctuation between the impeller and the stator.
Keywords: Axial-flow pump; Stator vane number; Numerical simulation; Performance test; Vibration; Pressure fluctuation
----------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------

1. Introduction
With high specific speed and outstanding capability of delivering working medium with large volume flow rate, the axialflow pump has been widely used in military, marine, agricultural and bio-medical fields. In most cases, axial-flow pumps
are installed vertically, which is prone to trigger adverse consequences such as vibration and failure of upper bearings [1]. In
addition, structural characteristics of the axial-flow pump are
complicated with the participation of turbulent flow. In recent
years, vibration sources nurtured in inner flows of the impeller
pump have gained much attention, and it has been verified in
some cases that static pressure fluctuations and pump vibration
share typical frequencies [2]. Most of these frequencies are
harmonics of shaft frequency or blade passing frequency.
The axial-flow pump is featured by unsteady flows between
impeller blades and stator vanes. Although numerous efforts
have been dedicated to the impellerstator interaction in the
axial-flow pump, a clear explanation of this subject is still
sorely necessitated [3]. Furthermore, issues like transient fluid
forces exerted on pump components have also been ascribed
*

Corresponding author. Tel.: +86 511 88780217, Fax.: +86 511 88780216
E-mail address: kangcan@ujs.edu.cn

Recommended by Associate Editor Donghyun You


KSME & Springer 2015

to the periodic interaction between rotating impeller blades


and stationary stator vanes [4]. This is because the working
medium discharged from the impeller has been energized and
possesses a large portion of total fluid kinetic energy. The
mixing of this part of fluid and adjacent fluid with a relatively
low level of fluid kinetic energy is conceivably intense [5]. In
this context, stator vane number plays an apparently important
role with its immediate effect on the periodicity of impellerrotor interaction. Glich investigated various combinations of
impeller blade number and stator vane number and proposed a
formula relating these two parameters [6].
For axial-flow gas turbines and compressors, another two
important branches of impeller machinery, rotor-stator interaction has been investigated extensively [7, 8]. Explicit pressure
spectra obtained in relevant experiments indicate that characteristic frequencies in the flow field between the rotor and the
stator correspond well to the rotational frequencies of impeller
blades. In contrast, for the axial-flow pump, inconvenience is
triggered by the liquid medium instead of gas medium. Thus
the design of a suitable test rig often makes many experimental schemes difficult to implement, so generalizable conclusions have rarely been reported hitherto [9]. For instance, optical measurement of the flow between impeller blades and
stator vanes seems impracticable in view of the demanding

2026

C. Kang et al. / Journal of Mechanical Science and Technology 29 (5) (2015) 2025~2034

configuration of optical transmitters and receivers [10]. With


vertical installation of the axial-flow pump, such a measurement is even questionable to accomplish [11]. Under such a
circumstance, pressure fluctuations, which could be obtained
through pressure transducers mounted on the pump casing, are
frequently used to describe flow characteristics.
Details of turbulent flows in the axial-flow pump may also
be revealed by using numerical simulation, an aggressively
popular technique which has been used to disclose diverse
flow structures inside impeller pumps. Years of experience
enable considerable advancement in both computation ability
and the reliability of numerical results. With numerical simulation, the accuracy of resultant averaged flow parameters and
hydraulic losses is ensured. Nevertheless, for time-dependent
parameters such as fluctuating velocity and pressure obtained
numerically, experimental validation is still badly needed. In
some cases, even with the same set of boundary and initial
conditions, different commercial computational fluid dynamics (CFD) codes might yield obviously different unsteady flow
patterns [12]. In essence, spatial and temporal scales required
for illustrating the instantaneous effect of flow structures on
pump vibration are exceedingly small, which often discourages the application of such an instrument.
Our major concern is the effect of stator vane number on
pump performance. Pump performance considered here incorporates inner flow feature, working capability of the pump,
pump vibration and static pressure fluctuation. To probe into
such a subject, numerical simulation is used as a preliminary
technique in seeking flow features immediately downstream
of the impeller and the stator as well. For comparison, pump
head and pump efficiency are measured for cases incorporating different vane numbers. Pump vibration is examined with
vibration acceleration sensors attached to both the supporting
plate of the pump and the motor base. High-frequency dynamic pressure sensors are used to measure static pressure
fluctuations between the impeller and the stator. And the results will account for how the local flow reacts to the periodically changed solid boundaries. Both flow and pump vibration
characteristics are supposed to be demonstrated as vane number varies. This study is anticipated to provide a further insight
into axial-flow pump performance, as well as inner flow characteristics in the pump with different combinations of impeller
blades and stator vanes.

2. Numerical model
2.1 Geometrical parameters and computational domain
Under nominal operation condition, the volume flow rate Q
of the axial-flow pump considered is 840 m3/h, and the corresponding pump head H is 5.2 m. The impeller spins at the
rotational speed n of 1470 rpm. Therefore, the value of specific speed ns of the pump is 206 based on the definition
ns =

n Q / 3600
.
H 0.75

(1)

Table 1. Primary geometrical parameters of the axial-flow pump.


Inlet diameter (mm)

300

Outlet diameter (mm)

300

Impeller diameter (mm)

300

Hub diameter (mm)

150

Impeller blade number Z

Stator vane number Z1

5, 7, 9

(a)

(b)

Fig. 1. Three-dimensional models for the 7-vane case: (a) hydraulic


components; (b) major computational subdomains.

Generally, empirical approaches are used in the design of


the impeller pump to yield preliminary results which are then
validated or optimized using CFD or experimental techniques
[13]. The streamline method was adopted to design the pump
impeller and the three stators. With such a method, variation
of circulation from the impeller hub to blade tip can be geared
towards an optimal distribution. Stator vanes were designed
based upon impeller blade parameters [14]. Primary geometrical parameters of hydraulic components are listed in Table 1.
All the three vane numbers-5, 7 and 9-conform to the general
knowledge of vane number selection. Most treatises relevant
to this procedure only suggest a consideration of excessive
frictional loss as large vane number is employed, but detailed
comparisons in vane number have rarely been documented
[15]. Three-dimensional geometrical models of the hydraulic
components associated with the 7-vane case are presented in
Figs. 1(a) and (b). The subdomains shown in Fig. 1(b) do not
include an auxiliary outlet straight subdomain, which is indispensable for full development of turbulent flow.
Unstructured grids were employed to discretize computational subdomains associated with impeller blade passages
and stator vane passages. Such a grid type can well accommodate irregular solid surfaces such as impeller blade and stator
vane surfaces. To improve the accuracy of numerical simulation and to detect small-scale flow phenomena residing in
local flows, grid refinement was executed for impeller and
stator computational subdomains. In line with the commonly
adopted approach evaluating the effect of grid number, a grid-

2026

C. Kang et al. / Journal of Mechanical Science and Technology 29 (5) (2015) 2025~2034

configuration of optical transmitters and receivers [10]. With


vertical installation of the axial-flow pump, such a measurement is even questionable to accomplish [11]. Under such a
circumstance, pressure fluctuations, which could be obtained
through pressure transducers mounted on the pump casing, are
frequently used to describe flow characteristics.
Details of turbulent flows in the axial-flow pump may also
be revealed by using numerical simulation, an aggressively
popular technique which has been used to disclose diverse
flow structures inside impeller pumps. Years of experience
enable considerable advancement in both computation ability
and the reliability of numerical results. With numerical simulation, the accuracy of resultant averaged flow parameters and
hydraulic losses is ensured. Nevertheless, for time-dependent
parameters such as fluctuating velocity and pressure obtained
numerically, experimental validation is still badly needed. In
some cases, even with the same set of boundary and initial
conditions, different commercial computational fluid dynamics (CFD) codes might yield obviously different unsteady flow
patterns [12]. In essence, spatial and temporal scales required
for illustrating the instantaneous effect of flow structures on
pump vibration are exceedingly small, which often discourages the application of such an instrument.
Our major concern is the effect of stator vane number on
pump performance. Pump performance considered here incorporates inner flow feature, working capability of the pump,
pump vibration and static pressure fluctuation. To probe into
such a subject, numerical simulation is used as a preliminary
technique in seeking flow features immediately downstream
of the impeller and the stator as well. For comparison, pump
head and pump efficiency are measured for cases incorporating different vane numbers. Pump vibration is examined with
vibration acceleration sensors attached to both the supporting
plate of the pump and the motor base. High-frequency dynamic pressure sensors are used to measure static pressure
fluctuations between the impeller and the stator. And the results will account for how the local flow reacts to the periodically changed solid boundaries. Both flow and pump vibration
characteristics are supposed to be demonstrated as vane number varies. This study is anticipated to provide a further insight
into axial-flow pump performance, as well as inner flow characteristics in the pump with different combinations of impeller
blades and stator vanes.

2. Numerical model
2.1 Geometrical parameters and computational domain
Under nominal operation condition, the volume flow rate Q
of the axial-flow pump considered is 840 m3/h, and the corresponding pump head H is 5.2 m. The impeller spins at the
rotational speed n of 1470 rpm. Therefore, the value of specific speed ns of the pump is 206 based on the definition
ns =

n Q / 3600
.
H 0.75

(1)

Table 1. Primary geometrical parameters of the axial-flow pump.


Inlet diameter (mm)

300

Outlet diameter (mm)

300

Impeller diameter (mm)

300

Hub diameter (mm)

150

Impeller blade number Z

Stator vane number Z1

5, 7, 9

(a)

(b)

Fig. 1. Three-dimensional models for the 7-vane case: (a) hydraulic


components; (b) major computational subdomains.

Generally, empirical approaches are used in the design of


the impeller pump to yield preliminary results which are then
validated or optimized using CFD or experimental techniques
[13]. The streamline method was adopted to design the pump
impeller and the three stators. With such a method, variation
of circulation from the impeller hub to blade tip can be geared
towards an optimal distribution. Stator vanes were designed
based upon impeller blade parameters [14]. Primary geometrical parameters of hydraulic components are listed in Table 1.
All the three vane numbers-5, 7 and 9-conform to the general
knowledge of vane number selection. Most treatises relevant
to this procedure only suggest a consideration of excessive
frictional loss as large vane number is employed, but detailed
comparisons in vane number have rarely been documented
[15]. Three-dimensional geometrical models of the hydraulic
components associated with the 7-vane case are presented in
Figs. 1(a) and (b). The subdomains shown in Fig. 1(b) do not
include an auxiliary outlet straight subdomain, which is indispensable for full development of turbulent flow.
Unstructured grids were employed to discretize computational subdomains associated with impeller blade passages
and stator vane passages. Such a grid type can well accommodate irregular solid surfaces such as impeller blade and stator
vane surfaces. To improve the accuracy of numerical simulation and to detect small-scale flow phenomena residing in
local flows, grid refinement was executed for impeller and
stator computational subdomains. In line with the commonly
adopted approach evaluating the effect of grid number, a grid-

C. Kang et al. / Journal of Mechanical Science and Technology 29 (5) (2015) 2025~2034

2027

production term connected with viscous force.


The other constants involved in Eqs. (2)-(4) are:
Cm = 0.0845 Ce 1 = 1.42 Ce 2 = 1.68

s k = 0.72 s e = 0.75.

In general, in flow regions with low strain rates, h < h0 ,


eddy viscosity calculated with the RNG k- model is higher
than that obtained with the standard k- model, and this tendency is overturned in high-strain-rate flow regions signified
by h > h0 . In the presence of both high strain rate and largecurvature solid walls, advantages of the RNG k- model are
prominent.
Fig. 2. Variation of pump head with grid number.

2.3 Discretization strategy and definition of boundary conditions

Dk
=
Dt xi

m
m + t
sk

k
+ Gk - re ,

xi

(2)

The commercial CFD code ANSYS-CFX served as a basic


platform for the numerical work covered in the present study
[18]. Central finite difference scheme was adopted to treat the
convection terms involved. Discretization of momentum and
turbulent kinetic energy equations is accomplished using
second-order upwind scheme. Velocity inlet boundary conditions were set at the inlet of the whole computational domain
and thus the magnitude of flow rate could be adjusted accordingly. The opening boundary condition in ANSYS-CFX was
set at the outlet of the whole computational domain. Such a
boundary condition is used when flow parameter distributions at the outlet of the whole computational domain are not
well understood. As working medium flows out of the computational domain, pressure is treated as static pressure.
While recirculation occurs there, pressure is treated as total
pressure. Non-slip condition was applied for all solid boundaries. Scalable wall functions were defined in near-wall flow
regions. At the interface between the impeller and the stator,
the frozen-rotor interface condition was specified. Such an
interface condition is a reliable steady-state strategy with
fluxes changing frame while the relative orientation of relevant components keeps invariant.

De

=
Dt xi

m
m + t
se

e
e
e2
*
,
+ Ce 1 Gk - Ce 2 r

k
k
xi

(3)

3. Steady simulation results

(4)

Two positions, impeller outlet and stator outlet, suffer directly from the effect of vane number, so do local flows involved. In view of currently attainable accuracy of numerical
simulation, only averaged flow information is extracted and
analyzed here. For the axial-flow pump, axial velocity component is of remarkable significance. Thus cross-sectional
distributions of axial velocity immediately downstream of the
impeller and the stator are constructed based upon the obtained numerical results. Relative velocity distributions at
impeller outlet are shown in Fig. 3 and absolute velocity distributions at stator outlet are shown in Fig. 4. In both figures,
positive velocity values denote the flow direction opposite to
the movement of main stream.
As indicated in Fig. 3, at a given vane number, the flow rate

independence examination was performed. Grid number was


determined after the difference of pump head between two
neighboring grid number schemes was less than 1%, as shown
in Fig. 2. Grid numbers involved in the final computation are
4671378, 5439416 and 6178812, which correspond to 5-vane,
7-vane and 9-vane cases, respectively.
2.2 Governing equations
The renormalization group (RNG) k- turbulence model,
first proposed by Yakhot and Orszag in 1986, indicates an
improvement relative to the standard k- turbulence model
[16]. This model embodies fuzzy mathematics principles, and
the parameters involved in this model are determined according to relevant formulae rather than empiricism or experiments. Along this line, the equation of turbulent kinetic energy
dissipation rate differs from its counterpart incorporated in
the standard k- turbulence model. The transport equations of
turbulent kinetic energy k and are:

and the variable Ce*2 in Eq. (3) is defined as

h
Cm rh 3 1 -
h
0

,
Ce*2 = Ce 2 +
1 + 0.012h 3

where dimensionless strain-rate parameter h is obtained


from h = Sk / e , and S is mean strain rate. h0 is a critical
value of dimensionless strain rate and h0 = 4.38 is predefined
[17]. Additionally, m and mt are viscosity and turbulent
viscosity, respectively. Gk denotes turbulent kinetic energy

2028

C. Kang et al. / Journal of Mechanical Science and Technology 29 (5) (2015) 2025~2034

Z1 = 5

Z1 = 7

Z1 = 9

(a) Flow rate of 0.8Q

Z1 = 5

Z1 = 7

Z1 = 9

(b) Flow rate of 1.0Q


Fig. 3. Distributions of relative velocity at impeller outlet.

of 0.8Q is associated with apparent reverse flow areas near the


impeller hub compared with the condition of nominal flow
rate. Meanwhile, the periodicity pertinent to the four impeller
blades is well maintained as vane number varies, particularly
near the impeller hub. Furthermore, the effect of stator vanes
can be perceived near the pump casing and is apparent at flow
rate of 0.8Q. As for the 9-vane case, the contribution of stator
vanes is obvious at both 0.8Q and 1.0Q. It is therefore deduced that large vane number leads to uniform cross-sectional
velocity distributions. Previous results obtained with particle
image velocimetry (PIV) also showed that from the impeller
hub to the pump casing, axial velocity increases and overall
magnitude of axial velocity increases with flow rate. Additionally, with PIV, vortices and recirculation are revealed near
the impeller hub and vortices are especially distinct at small
flow rates [19].
At stator outlet, the energy transformation capability of stator vanes is well manifested, as shown in Fig. 4. On the whole,
the uniformity of axial velocity distribution is enhanced as
vane number increases. Furthermore, as vane number in-

creases, reverse-flow areas near the stator hub tend to be annihilated. Flow rate also plays an important part in this connection. An increase in flow rate restricts the development of
reverse flows near the pump casing, as applies to the three
cases. The most favorable situation surfaces at vane number of
9 and a corresponding flow rate of 1.0Q, while the most adverse situation is evidently related to the 5-vane case at flow
rate of 0.8Q. For the latter, the existence of large-scale reverse
flow structures which occupy a large portion of cross-sectional
area evidently reflects the deficiency of energy transformation.

4. Comparison of operation performance among three


pumps
Several studies confirmed the connection between flow
characteristics and operation performance of the pump, but no
robust explanation has been furnished so far [20]. Since some
factors such as mechanical loss and leakage loss were not
taken into account in numerical simulation, the operation performance of the pump predicted numerically inevitably in-

2029

C. Kang et al. / Journal of Mechanical Science and Technology 29 (5) (2015) 2025~2034

Z1 = 5

Z1 = 7

Z1 = 9

(a) Flow rate of 0.8Q

Z1 = 5

Z1 = 7

Z1 = 9

(b) Flow rate of 1.0Q


Fig. 4. Absolute velocity distributions at stator outlet.

(a) Variation of pump head with flow rate

(b) Variation of pump efficiency with flow rate

Fig. 5. Operation performance of three axial-flow pumps.

corporated errors. Under such a circumstance, a performance


test was carried out on the test platform in Shanghai Kaiquan
Pump Group Co., Ltd. Through adjustments and regulations
performed on the experimental system, the maximum uncer-

tainties for measurements of pump head and pump efficiency


were less than 1% and 1.8%, respectively. Both pump head
and pump efficiency were measured for the three pumps with
different vane numbers and the results are plotted in Fig. 5.

2030

C. Kang et al. / Journal of Mechanical Science and Technology 29 (5) (2015) 2025~2034

(a) Arrangement of monitored points, P1-P4 on the motor base

(c) An overall view of the test rig

(b) Arrangement of monitored points, P1-P4,on the supporting plate of the pump

(d) Sketch and image of the pressure hole for pressure fluctuation measurement

Fig. 6. Experimental apparatus for measurements of vibration and pressure fluctuation.

As for the variation of pump head with flow rate shown


in Fig. 5(a), it is observable that the three curves do not
deviate from each other obviously, although three significantly different sets of stators are being used. Note that for
the three cases, a saddle-shape pump head curve segment, a
distinct feature of the axial-flow pump, is unanimously
found near flow rate of 500 m3/h, namely 0.6Q. Since the
three impellers are identical, this experimental result also
supports the acknowledged conclusion that reverse flows
and circulation near impeller inlet facilitate the saddleshape performance [21].
In regard to the relationship between pump efficiency and
flow rate, as illustrated in Fig. 5(b), the 7-vane pump holds the
highest efficiency, which corresponds to a flow rate slightly
higher than the nominal flow rate. As for the 9-vane pump, the
best efficiency point is situated far from the nominal operation
condition, and the overall pump efficiency of the 9-vane pump
is the lowest among the three pumps. According to the previous discussion, large vane number helps to regulate the flow
issuing from the impeller. However, the increase of friction
loss occurring concurrently might in turn induce a decline in
pump efficiency.

5. Pump vibration and static pressure fluctuations


5.1 Experimental methodology
The importance of pump vibration is increasingly being
recognized, as is partially attributed to advancement in relevant research techniques [22]. As a subject stressed here, the
vibration of the axial-flow pump is analyzed from the perspective of experiment instead of numerical simulation. In the light
of conventional pump design principles, pump vibration of a
minimum degree is attained at nominal flow rate. Nevertheless, no quantitative verification has been demonstrated even
at some given specific speed [23]. Consequently, the contribution of flow parameter distribution or flow pattern to pump
vibration has not been substantiated. In the present study, an
LMS multi-channel vibration and noise measurement system
was used to measure both pump vibration and motor vibration.
Sensors for vertical vibration acceleration measurement were
attached to specified local surfaces. Monitored points on the
motor base, P1 to P4, are shown in Fig. 6(a) and these four
points are located on the same horizontal plane. Another four
coplanar points, P1 to P4, are deployed on the supporting plate
of the pump, as shown in Fig. 6(b). The pump equipped with

C. Kang et al. / Journal of Mechanical Science and Technology 29 (5) (2015) 2025~2034

Fig. 7. Vibration acceleration level measured on the motor base at


nominal flow rate.

2031

(a) 5-vane case

the 7-vane stator is shown in Fig. 6(c) which also shows the 5vane and 9-vane stators. Aside from overall vibration, transient static pressure fluctuations were measured with several
miniature dynamic pressure sensors mounted circumferentially on the pump casing, and these sensors were axially located in the middle of the impeller and the stator. Meanwhile,
these sensors were fitted into pressure holes and sensor tips
protruded into working medium. One of the pressure holes is
shown in Fig. 6(d). The frequency range of 10-8000 Hz was
predefined for data acquisition.
5.2 Pump vibration
Transiently captured data with the sensors were processed
using fast Fourier transformation (FFT) technique, converting
time-dependent data into signals in frequency domain. Due to
fluctuations of the rotational speed of the impeller due to
power frequency fluctuations of the electric power grid, the
actually monitored shaft frequency, f, is 24.61 Hz. There is a
relative deviation of 1.24% compared with the nominal shaft
frequency, which is acceptable as per requirements of ISO
9906:2012. Furthermore, the blade passing frequency (BPF),
fz, is 99.22 Hz, which is in accordance with the impeller blade
number of 4.
In view of the configuration of pump unit components, the
four monitored points deployed on the motor base present not
just the vibration of the motor itself but also the vibration of
the upper part of the pump. However, characteristic frequencies associated with this set of data hinge upon multiple factors and cannot be conveniently predicted. Through data processing, maximum and average values of vibration acceleration
level at nominal flow rate are obtained and the results are plotted in Fig. 7.
In Fig. 7, for each monitored point, particularly P1, a large
gap exists between average value and corresponding maximum value. Such a situation implies that local vibration endures transient fluctuations. Nevertheless, in an average sense,
vibration levels at the four monitored points are nearly identi-

(b) 7-vane case

(c) 9-vane case


Fig. 8. Vibration acceleration level measured on the supporting plate of
the pump.

cal, irrespective of the difference in vane number. It should be


stressed that the monitored point P1 is adjacent to the outlet
bend of the pump, and the effect of the bend on local vibration
should be taken into account. In this context, the deflection of
the working medium in the bend and the structure of the bend
jointly constitute an influential factor aggravating local vibration. By contrast, the overall vibration level associated with P3

2032

C. Kang et al. / Journal of Mechanical Science and Technology 29 (5) (2015) 2025~2034

is comparatively low. As for another two monitored points, P2


and P4, the overall vibration levels well prove the connection
between vibration and local position. In addition, Fig. 7 indicates that the influence of vane number on average vibration
acceleration level is not remarkable, but the maximum vibration acceleration levels corresponding to the 7-vane case are
low compared with their counterparts.
As flow rate increased from 0.8Q to 1.2Q with an increment
of 0.2Q, vibration acceleration level was measured on the
supporting plate of the pump, and the results are plotted in Fig.
8. For all the three cases, the vibration acceleration level values at monitored points P2 and P4 are quite close. In this connection, the line covering P2 and P4 is perpendicular to the
inflow and outflow straight pipes connected to the pump.
Therefore, it can be concluded that circumferential balance of
the pump is preferable without obvious lateral vibration.
Nominal flow rate is associated with minimum vibration, as
applies to monitored points P1 and P3, although vibration
levels at these two points are particularly high. In view of the
positions of P1 and P3, characteristics of the entire system in
which the pump operates contribute apparently to the local
vibration considered. In particular, the monitored point P3,
with the highest vibration level, is situated in a diagonal position with respect to the monitored point P1, as shown in Figs.
6(a) and (b). And P3 is also the nearest point from the inflow
pipe. The three cases share similar overall variation tendency
of vibration as flow rate increases. And the highest vibration
level occurs at 0.8Q. Furthermore, for each case, although
vibration acceleration levels at P2 and P4 differ, they are not
suitable for assessing the influence of flow on vibration acceleration level. Note that at vane number of 9, at P2 and P4,
variation of vibration acceleration level with flow rate is comparatively regular.
5.3 Pressure fluctuations at 0.8Q
As a further exploration, pressure fluctuation between the
impeller and the stator is analyzed. The contribution of turbulent fluctuation to the vibration of the whole pump has been
highlighted in recent studies [24]. The experimental system
used in this study enables the measurement of time-dependent
static pressure fluctuations with a high fidelity. In accordance
with the main aim of the present study, static pressure fluctuations in the flow region between the impeller and the stator
serve as an indicator of turbulent flow features, which are
sensitive to the effect of vane number. By using fast Fourier
transformation, pressure fluctuations in frequency domain at
flow rate of 0.8Q were obtained and plotted in Fig. 9.
For all the three cases, blade passing frequency and its harmonics dominate the frequency spectra, as evidenced in Fig. 9.
In comparison, shaft frequency and its harmonics can only be
identified through some considerably low peaks of pressure
fluctuation amplitude. These transient pressure fluctuations
obtained demonstrate that the rotation of the impeller is highly
influential even at small time scales. Also shown in Fig. 9, for

Fig. 9. Pressure fluctuations between the impeller and the stator at 0.8Q.

all the three cases, high-frequency components within the


displayed frequency range are discretely distributed. In general, small-scale flow structures stimulate the emergence of
high-frequency components [25]. At high frequencies, the 9vane case incorporates high amplitudes of pressure fluctuations compared with the other two cases. This is inseparable
from the prevalence of small flow structures at vane number
of 9. In addition, these high frequencies are still multiples of
blade passing frequency. For each case, low peaks are present
among high pressure fluctuation amplitude peaks, as is particularly clear at vane number of 9.
The overall pressure fluctuation amplitude of the 7-vane
case is lower than its counterparts. Meanwhile, the similarity
between the frequency spectra of the 5-vane and 7-vane cases
is easily recognizable, while the frequency spectrum of the 9vane case is quite different. From another perspective, the
combination of five stator vanes and four impeller blades
boosts the formation of large-scale flow structures, which
offer little support to high-frequency pressure fluctuations.
Additionally, for the 9-vane case, periodic interaction between
impeller blades and stator vanes is frequent and intense, irrespective of the uniformity of velocity distribution. In Ref. [6],
it is suggested that large discrepancy in periodicity, such as the
situation of four impeller blades along with nine stator vanes,
enhances flow-induced pump vibration.
5.4 Pressure fluctuations at 1.0Q
Pressure fluctuations at nominal flow rate are illustrated in
Fig. 10. And the maximum pressure fluctuation amplitude
occurs invariably at blade passing frequency; overall pressure
fluctuations are appreciably suppressed compared with Fig. 9.
In regard to the 5-vane case, the two pressure fluctuation spectra corresponding to 0.8Q and 1.0Q are highly analogous.
Nevertheless, the pressure fluctuation amplitude at blade passing frequency decreases sharply from about 20 kPa to 12.5
kPa. Furthermore, discrete peaks of pressure fluctuations

C. Kang et al. / Journal of Mechanical Science and Technology 29 (5) (2015) 2025~2034

2033

cases. Compared with the other two cases, the 7-vane case has
relatively low characteristic pressure fluctuation amplitudes.
As flow rate increases from 0.8Q to 1.0Q, all characteristic
pressure fluctuation amplitudes decline with the 5-vane case,
while the 9-vane case is featured by active low-order harmonics together with decayed high-frequency components.

Acknowledgments

Fig. 10. Pressure fluctuations between the impeller and the stator at 1.0Q.

This study is financially supported by the project funded by


College Industrialization Project of Jiangsu Province (Grant
No.JHB2011-37) and Priority Academic Program Development of Jiangsu Higher Education Institutions. The authors
would extend their gratitude to all colleagues and graduate
students actively involved in the experimental work.

Nomenclature-----------------------------------------------------------------------among harmonics of blade passing frequency are maintained.


In contrast, the other two cases, especially the 9-vane case,
experience an overall change as flow rate increases. For the 7vane case, all pressure fluctuation amplitudes decrease as flow
rate increases, while the pressure fluctuation amplitude corresponding to blade passing frequency varies slightly. A distinct
characteristic pertinent to the 7-vane case is that most highfrequency components are restrained or even eliminated as
flow rate increases from 0.8Q to 1.0Q. Concerning the 9-vane
case, many high-frequency components arising at 0.8Q are
suppressed. Additionally, harmonic frequencies of fz are active
at vane number of 9, as cannot evade from the influence of
local flow patterns. This connection has not been detailed
since correlations among flow structures have not been established due to lack of turbulence data.

6. Conclusions
Vane number affects both energy transformation capability
and inner flow characteristics of the axial-flow pump. Pump
head curves associated with the three vane number cases are
in modest agreement, while the 7-vane case proves to be the
most preferable one in terms of pump efficiency. The 9-vane
case has the lowest overall pump efficiency. Nevertheless,
large vane number facilitates the rise in velocity distribution
uniformity at both impeller outlet and stator outlet.
Variation in vane number accounts for an insignificant
change in terms of pump vibration. Apart from the circumferentially even vibration monitored on the motor base, high
vibration level arises at two local positions on the supporting
plate: one is near the outlet bend of the pump and the other is
adjacent to the inflow pipe. Both structural and hydraulic factors are pivotal in this context. At nominal flow rate, minimum pump vibration level is manifested.
Frequency spectra of pressure fluctuations between the impeller and the stator are overwhelmingly dominated by blade
passing frequency and its harmonics, as is shared by the three

fz
H
k
n
ns
Q
Z
Z1

: Blade passing frequency


: Pump head
: Turbulent kinetic energy
: Rotational speed
: Specific speed
: Volume flow rate
: Impeller blade number
: Stator vane number
: Turbulent kinetic energy dissipation rate

References
[1] E. Egusquiza, C. Valero, X. Huang, E. Jou, A. Guardo and C.
Rodriguez, Failure investigation of a large pump-turbine
runner, Engineering Failure Analysis, 23 (2012) 27-34.
[2] M. Vahdati, A. I. Satma and M. Imregun, An integrated
nonlinear approach for turbomachinery forced response prediction, Part II: case studies, J. of Fluid and Structures, 14
(2000) 103-125.
[3] D. Zhang, W. Shi, T. Li, H. Zhang and X. Guan, Property of
unsteady pressure and meridional velocity in wake region of
axial-flow pump impeller, Transactions of the Chinese Society of Agricultural Engineering, 28 (17) (2012) 32-37.
[4] P. Wei, H. Chen and W. Lu, Characteristics of force acting
on adjustable axial flow pump blade, Frontiers of Energy
Power Engineering in China, 2 (4) (2008) 508-513.
[5] D. Kaya, Experimental study on regaining the tangential
velocity energy of axial flow pump, Energy Conversion and
Management, 44 (2003) 1817-1829.
[6] J. F. Glich, Centrifugal pumps, Second Ed., Springer, Berlin Heidelberg, Germany (2010).
[7] X. J. Yu and B. Liu, Stereoscopic PIV measurement of unsteady flows in an axial compressor stage, Experimental
Thermal and Fluid Science, 31 (2007) 1049-1060.
[8] N. Courtiade, X. Ottavy and N. Gourdain, Modal decomposition for the analysis of the rotor-stator interactions in multistage compressors, J. of Thermal Science, 21 (3) (2012) 276-

2034

C. Kang et al. / Journal of Mechanical Science and Technology 29 (5) (2015) 2025~2034

285.
[9] H. Wang and H. Tsukamoto, Fundamental analysis on rotorstator interaction in a diffuser pump by vortex method, J. of
Fluids Engineering, 123 (2001) 737-747.
[10] E. Gttlich, J. Woisetschlger, P. Pieringer, B. Hampel and
F. Heitmeir, Investigation of vortex shedding and wakewake interaction in a transonic turbine stage using laserDoppler-velocimetry and particle-image-velocimetry, J. of
Turbomachinery, 128 (1) (2006) 178-187.
[11] Y. Wu, S. Liu, H. Yuan and J. Shao, PIV measurement on
internal instantaneous flows of a centrifugal pump, Science
China Technological Science, 54 (2011) 270-276.
[12] D. Zhang, W. Shi, B. Chen and X. Guan, Unsteady flow
analysis and experimental investigation of axial flow pump,
J. of Hydrodynamics (Ser. B), 22 (1) (2010) 35-43.
[13] S. Kim, K.-Y. Lee, J.-H. Kim, J.-H. Kim, U.-H. Jung and
Y.-S. Choi, High performance hydraulic design techniques
of mixed-flow pump impeller and diffuser, Journal of Mechanical Science and Technology, 29 (1) (2015) 227-240.
[14] C. Kang, Q. Huang and Y. Li, Influence of vane number on
the performance of axial-flow pump under low-flow-rate
conditions, Proc. of ASME FEDSM 2013, Nevada, USA
(2013) Doi:10.1115/FEDSM2013-16331.
[15] Z. Qian and Y. Wang, Numerical simulation of water flow
in an axial flow pump with adjustable guide vanes, JMST, 24
(4) (2010) 971-976.
[16] V. Yakhot and S. A. Orszag, Renormalization group analysis of turbulence, J. of Scientific Computing, 1 (1986) 3-51.
[17] Y. Wu, S. Liu, H.-S. Dou, S. Wu and T.n Chen, Numerical
prediction and similarity study of pressure fluctuation in a
prototype Kaplan turbine and the model turbine, Computers
& Fluids, 56 (2012) 128-142.
[18] ANSYS CFX-Solver Theory Guide, ANSYS Inc. (2009).
[19] H. Huang, H. Gao and Z. Du, Numerical simulation and

experimental study on flow field in an axial flow pump, J. of


Shanghai Jiaotong University, 43 (1) (2009) 124-128.
[20] H. Bing, S. Cao, C. He and L. Lu, Experimental study of
the effect of blade tip clearance and blade angle error on the
performance of mixed-flow pump, Science China Technological Science, 56 (2013) 293-298.
[21] Y. Zheng, F. Zhang, X. Jiang and J. Zhang, Research on
outlet pressure pulsation of reverse operation in tubular
pump model experiment, Fluid Machinery, 35 (1) (2007) 1-7.
[22] Z. Li, M. Yang, N. Zhang and B. Gao, Experimental study
on vibration characteristics of axial-flow pump under different operating conditions, J. of Engineering Thermophysics,
34 (5) (2013) 866-869.
[23] F. Shi and H. Tsukamoto, Numerical study of pressure
fluctuation caused by impeller-diffuser interaction in a diffuser pump stage, J. of Fluids Engineering, 123 (9) (2001)
466-474.
[24] C. Chen, A. Yang, G. Li, P. Zhou and Q. Wang, Fluid excitation forces of axial flow pumps with different number of
blades, Noise and Vibration Control, 33 (3) (2013) 55-59.
[25] Z. Li, M. Yang, B. Gao and C. Kang, Experimental study
on vibration characteristics induced by cavitation of axialflow pump, J. of Engineering Thermophysics, 33 (11) (2012)
1887-1891.

Can Kang received his Ph.D. from the


Jiangsu University of China. He is an
associate professor and Ph.D. supervisor
at the Jiangsu University of China. His
research interests include optimal design
of impeller pumps, non-intrusive flow
measurement, high-pressure water jet
and cavitation in aerated flows.

You might also like