Applied Mathematical Modelling: Wen-Guang Li
Applied Mathematical Modelling: Wen-Guang Li
a r t i c l e i n f o a b s t r a c t
Article history: The performance and flow of a low specific speed experimental centrifugal pump as turbine
Received 1 April 2014 are identified by means of CFD method when it handles a more viscous liquid than water. The
Revised 12 May 2015
performance curves are compared among five kinds of liquids. Meanwhile the internal vari-
Accepted 24 June 2015
ables, such as impeller theoretical head, total hydraulic loss coefficients across the machine,
Available online 11 July 2015
individual total loss coefficient in suction pipe/draft tube, impeller and volute, averaged skin
Keywords: friction factors, pressures coefficients over blade surfaces as well as incidence loss are ex-
Centrifugal pump plored. The correction factors of flow rate, head, hydraulic efficiency and impeller theoreti-
Pump as turbine cal head in both turbine and pump modes are correlated to impeller Reynolds number by a
Viscosity three-parameter-power function. The flow behaviours in turbine mode are characterized by
Reynolds number examining the velocity magnitude and flow angle at the volute outlet, angle of attack, fluid
Correction factor velocity vectors in the volute, swirling patterns of flow in the draft tube and deviation angle at
CFD impeller outlet. The results can be useful for pump model selection, performance prediction,
hydraulic design and impeller geometry modification.
© 2015 Elsevier Inc. All rights reserved.
1. Introduction
At present, a centrifugal pump can be operated in a reverse direction as a renewable energy generator in micro-hydro-power
plants [1–7] or as a power recovery turbine in the process industry, such as reverse osmosis, hydrocarbon process, gas scrubbing
process, ammonia production, petroleum cracking process, hot water circulation system and so on [8–10]. The application of
pump as turbine (PAT) is very helpful to reduction of carbon dioxide emission and maintenance of sustainable development
of human being. Against normal hydro-turbines, a centrifugal pump as turbine is cheaper and more feasible and available for
choice.
In order to select a proper centrifugal pump for a turbine application, we need to know how to predict the turbine perfor-
mance based on a pump performance available. Therefore, various approaches for estimating turbine performance have been
proposed [1,11]. Besides, a series of experiments on performance and flow of PAT have been conducted in [12–20] since the
1990s. Particularly, experimental performance curves of a pump as turbine were obtained at different specific speeds in [12,13]
by Derakhshan et al. An optimization approach for performance prediction and selection of a pump as turbine was proposed
by Singh and Nestmann in [14]. The performance of a pump as turbine and flow field up- and down-stream the impeller were
measured by Shinhama et al. in [18] and by Fernandez et al. in [20]. These respected studies have improved our understanding
about PAT performance and inside flow features.
∗
Tel.: +86 931 2756251; fax: +86 931 2576250.
E-mail address: [email protected]
http://dx.doi.org/10.1016/j.apm.2015.06.015
S0307-904X(15)00393-5/© 2015 Elsevier Inc. All rights reserved.
W.-G. Li / Applied Mathematical Modelling 40 (2016) 904–926 905
Nomenclature
Greek letters
α fluid absolute flow angle measured from a meridian plane, deg
β fluid relative flow angle measured from a meridian plane, deg
βb blade angle measured from a meridian plane, deg
β angle of attack in impeller inlet or deviation angle at impeller outlet in turbine mode, deg
φ flow deviation angle from a blade at leading edge due to the slip effect in turbine mode, estimated by using the
Stodola slip factor formula, deg
ε rate of dissipation of turbulence kinetic energy per unit mass due to viscous stresses, m2 /s3
ηh hydraulic efficiency
κ turbulence constant in Eq. (5)
λ total skin friction factor in a hydraulic component, calculated from the total friction drag on the wet walls, mean
fluid velocity and mass flow rate
μ dynamic viscosity of liquid, Pa s
μt Turbulence eddy viscosity, Pa s
ν fluid kinematic viscosity, cSt (mm2 /s)
ξ total hydraulic loss coefficient in pump and turbine modes or incidence loss coefficient in turbine mode
ρ fluid density, kg/m3
σk , σε constants in the k– ε turbulence model
τ average shear stress on wet wall, Pa
ϕ helix angle, deg
ω rotational angular speed of impeller, rad/s
906 W.-G. Li / Applied Mathematical Modelling 40 (2016) 904–926
−
→
ω rotational angular speed vector of impeller, rad/s
Subscripts
0 volute outlet in turbine mode
1 impeller inlet in pump or turbine mode
2 impeller outlet in pump or turbine mode
e impeller theoretical head
H head
i impeller
in incidence
min minimum value
p pump mode
Q flow rate
s suction pipe or draft tube
t turbine mode
a axial direction
u circumferential direction
v volute
w water
η hydraulic efficiency
Recently, CFD methods start to be applied in PAT performance prediction and flow pattern characterization. In [21-24], steady
turbulent flows in PAT were simulated by using various CFD codes and the performance was established based on the simulated
results, especially those by Gonzalez et al. in [22] and by Nautiyal et al. in [24]. Unsteady flow fields in PAT were tackled by means
of moving mesh technique in CFD codes by Morros et al. in [25] and by Barrio et al. in [26]. The PAT transient characteristics
were clarified numerically by Fecarotta et al. in [27] as well. The steady and unsteady radial thrusts are estimated by using CFD
simulations by Fernandez et al. in [20], by Gonzales et al. in [22] and Couzinet et al. in [28], showing a PAT can experience an
increased radial thrust compared with the pump mode.
In particular, CFD and experimental methods have been applied to identify an improvement in performance and flow pattern
when there is a change in leading edge shape of PAT impeller blade by Singh and Nestmann in [29], or when the PAT impeller
diameter is cut down by Yang et al. in [30] or when the PAT guide vane installation pattern is altered by Patel et al. in [31]. These
fruitful outcomes suggest CFD methods have increasingly become an important tool for investigation into PAT.
Note that in those PAT studies water is the only fluid used. This does not seem true in PAT application, namely in the petroleum
cracking process where the fluid is often more viscous than water. How the liquid viscosity affects PAT performance and flow
pattern is open for questions. This problem has not appeared to be tackled so far.
Here we investigate the performance and flow fields of a PAT, which is a low specific speed experimental centrifugal pump as
turbine at the same rotational speed as the pump, by using CFD method at various viscosities. The flow rate, head and hydraulic
efficiency correction factors in either pump or turbine mode have been examined and correlated to impeller Reynolds number by
using a three-parameter-power function. The flow details in turbine mode are analysed and highlighted. It is confirmed that the
liquid viscosity influences PAT performance more pronouncedly than does the pump performance, the PAT impeller is subject to
a reverse flow and the draft tube involves a swirling flow at any operating points, and the increased hydraulic losses in impeller
and draft tube attribute to reduced turbine performance at a higher viscosity.
The computational model is the single-stage, end-suction experimental centrifugal pump in [32]. At the design point, the
pump flow rate, Q p , head, H p , and rotational speed, n p , are 25 m3 /h, 8 m and 1450 r/min, and its specific speed is defined as
nsp = 3.65n p Q p /H p3/4 = 93 (r/min, m3 /s, m). The impeller eye and outlet diameters are 62 mm and 180 mm, the number of
blades is 4, the exit blade angle is 70° (the angle between blade camber line and the meridian plane), the warp angle of blade is
140°. The diameter of base circle of the volute is 190 mm, the volute width is 40 mm, and the cross-sectional area of its throat is
1440 mm2 .
The computational fluid domains are composed of the stationary suction pipe, rotating impeller and unmovable volute, see
Fig. 1(a) and (b). Interface 1 and 2 are generated by linking the suction pipe outlet and the impeller entrance as well as by
matching the impeller outlet and the volute inlet, respectively. The coupling of flow variables on the interfaces is fulfilled by
means of the multiply reference frame (MRF) method in FLUENT6.2.
The fluid domains and velocity triangles at the inlet and outlet of the impeller are illustrated in Fig. 1(c) and (d) in pump
and turbine modes. In the pump mode, the impeller is under a counterclockwise rotation when it is viewed from the suction
side against the impeller entrance. At the design or best efficiency point (BEP), a liquid is sucked into the impeller without any
pre-swirl; then the tangential component of absolute velocity of the liquid is raised, but also the liquid static pressure becomes
W.-G. Li / Applied Mathematical Modelling 40 (2016) 904–926 907
Fig. 1. Computational fluid domains and velocity triangles in pump and turbine modes as well as mesh structure, (a) pump fluid domain, (b) turbine fluid
domain, (c) inlet and outlet velocity triangles in pump mode, (d) inlet and outlet velocity triangles in turbine mode, (e) mesh structure through the mid-span
plane of impeller, (f) mesh structure in horizontal plane through the shaft axis.
higher because of the centrifugal force effect and deceased relative velocity in the impeller under the action generated by the
rotating blades; in consequence, a head rise is developed in the liquid.
In the turbine mode, however, the liquid exhibits a clockwise circulation at the exit of volute and impacts the blades to put
the impeller into rotation and generate an output shaft power. The liquid discharged from the impeller possesses a reduced
tangential component of absolute velocity and lowered static pressure; eventually it is discharged out of the turbine from the
draft tube. Note that the absolute flow angle in Fig. 1(c) and (d) is the angle between the absolute velocity and the meridian
plane in the impeller; similarly, the relative flow angle is the angle between the relative velocity and the meridian plane as well,
see [33].
The mesh structure and number of cells are identical to those in [32] and illustrated in Fig. 1(e) and (f). This is a hybrid mesh
with Cartesian cells inside the core of the fluid domain and tetrahedral cells close to the boundaries generated in GAMBIT. In the
908 W.-G. Li / Applied Mathematical Modelling 40 (2016) 904–926
fluid domain of suction pipe, the number of hexahedral cells is 89,000, but in the impeller and volute fluid domains, the numbers
of hybrid cells are 231,860 and 351,201. This set of mesh is referred to as fair mesh.
To check if the performance parameters of turbine in CFD is independent of mesh size, a fine mesh was created as well, namely
200,431 hexahedral cells in the draft tube and 491,070 hybrid cells in the impeller and 543,675 hybrid cells in the volute. This set
of mesh is named as fine mesh.
It is shown that the head and hydraulic efficiency with the fair mesh differ from those with fine mesh by ±3% and ±2% errors
only in both pump and turbine modes at BEP, suggesting the fair mesh is accurate enough for fluid flow simulations. All the
results presented in the following sections are based on the fair mesh.
The fluid is incompressible, and its flow is 3D and turbulent inside the pump or turbine under any operation conditions.
The time-averaged flow of the fluid is steady. The fluid flow is governed by the time-averaged continuity and Navier–Stokes
equations. In the MRF system in [34], the continuity equation of fluid flow is
∂ρ −
→ − → −→
+ ∇ · [ρ( V − ω × r )] = 0 (1)
∂t
And the momentum equations of the fluid flow in pump and turbine modes can be expressed by
∂ −
→ → − −
→ −
→ −
→T
ω ×→
(ρV ) + ∇ · [ρ( V − − r ) V ] + ρ(ω
−
→
× r ) = −∇ p + ∇ · [(μ + μt )(∇ V + ∇ V )] (2)
∂t
→ −
− → −
→ −
→
where ω = e z ω in the impeller of pump mode, but ω = − e z ω in the turbine mode. In a stationary component, however,
−
→
ω = 0 in both modes.
The standard k − ε two equation turbulence model is selected to estimate the turbulence eddy viscosity, μt . The k and ε
equations are as follows [34]
∂ −
→
μt
(ρ k) + ∇ · ρ V k = μ + ∇ k + Gk − ρε (3)
∂t σk
and
−
∂ → μt ε
(ρε) + ∇ · ρ V ε = μ+ ∇ε + (C1ε Gk − C2ε ρε) (4)
∂t ση k
where the turbulent viscosity, μt = ρCμ k2 /ε , the production of turbulence kinetic energy k, Gk is computed from Gk =
−
→ −
→ T
μt (∇ V + (∇ V ) ) : ∇V . The model constants C1ε , C2ε , Cμ , σk and σε have the default values, such as 1.44, 1.92, 0.09, 1.0 and
1.3, respectively [34].
The non-equilibrium wall function is applied to involve wall shear stress as to account for the effect of pressure gradient in
the primary flow direction on the wall shear stress, which is given in [34] as follows
⎧ 1/4 1/2
⎪ ŨCμ k 1 ρCμ1/4 k1/2 y
⎪
⎨ τ /ρ = ln E
κ μ
y y−y
(5)
⎪
⎪ y2c
⎩Ũ = U − 1 dp yc
ln +
c
+
2 dL ρκ k1/2 yc ρκ k1/2 μ
Clearly, dp/dL effect on the wall shear stress τ is involved in Eq. (5).
The finite volume method is chosen to discretize the fluid flow governing equations. The continuity and momentum equations
are coupled with SIMPLE. The pressure in the momentum equation is interpolated by using the staggered scheme. The convection
items in the momentum, k and ε equations are discretized in the 2nd order upwind scheme, but the diffusing items in them are
done all in the 2nd order central differential formula.
The under relaxation factors are 0.2, 0.5, 0.8 and 0.8 for the continuity (pressure correction), momentum, k and ε equations,
respectively. The convergence criterion of the residual in those equations is 1 × 10−4 .
In the pump mode, a velocity-inlet boundary condition is applied at the suction pipe entrance, and the normal velocity mag-
nitude is calculated with the flow rate specified and the cross-sectional area of the pipe. A pressure-outlet boundary condition is
imposed at the pump (volute) outlet, the pressure magnitude is 0 Pa.
In the turbine mode, there is a velocity-inlet boundary condition at the turbine inlet (volute outlet), the normal velocity
magnitude is estimated with the flow rate specified and the cross-sectional area. There is a pressure-outlet boundary condition
at the outlet of draft tube (the entrance of suction pipe) where the pressure level is 0 Pa.
The rest surface is wet wall and hydraulic smooth in both modes. The effects of wall roughness on performance and flow field
of a turbine will be clarified in other investigations.
The working fluids in the simulations are water and four kinds of viscous oils, respectively, their kinematic viscosity and
density as well as impeller Reynolds number at 20 °C are tabulated in Table 1. The impeller Reynolds number is defined as
Re = u1t r1t /ν = u2p r2p /ν .
W.-G. Li / Applied Mathematical Modelling 40 (2016) 904–926 909
Table 1
Physical parameters of working fluids at 20 °C and impeller Reynolds number.
Liquid Density, ρ (kg/m3 ) Kinematic viscosity, ν (mm2 /s) Dynamic viscosity, μ(Pa s) Impeller Reynolds number, Re
Fig. 2. Head, shaft-power and hydraulic efficiency of pump and turbine are plotted as a function of flow rate, (a) head, (b) shaft-power, (c) hydraulic efficiency.
3. Results
The effects of liquid viscosity on the head, output shaft power and hydraulic efficiency curves of turbine mode are demon-
strated in Fig. 2 as a function of turbine flow rate. For comparison, the head, shaft power, and hydraulic efficiency curves of pump
are included in terms of pump flow rate, too. The head consumed by the turbine is always higher than the head generated by the
pump; further the former rises with increasing flow rate, exhibiting a positive slope and its magnitude is larger than the absolute
value of a slope of the latter at any viscosities.
This implies that the flow pattern in the turbine is unsatisfactory compared with the pump. Note that the higher the liquid
viscosity, the more the head consumed by the turbine, especially a flow rate lower than 13 L/s.
The turbine does not generate useful shaft power until the flow rate of 7 L/s, but also the turbine shaft power does not exceed
the pump until the flow rate of 15 L/s. If a turbine is wanted to develop more shaft power, a higher flow rate should be allowed to
pass by it. Similar to the head, the turbine shaft power ascends with the increasing flow rate; also its slope is obviously greater
than the pump shaft power.
Since the density of water is higher than the viscous oils, the turbine generates more shaft power with water than with any
viscous oils. Particularly, the density of the four kinds of viscous oils is quite close to each other, thus the power generated with
them is nearly identical.
910 W.-G. Li / Applied Mathematical Modelling 40 (2016) 904–926
Fig. 3. Moment of fluid absolute velocity at impeller inlet and outlet, impeller theoretical head in terms of flow rate in pump and turbine modes at five viscosities,
(a) moment at impeller inlet in pump mode and at outlet in turbine mode, (b) moment at impeller outlet in pump mode and at inlet in turbine mode, (c) impeller
theoretical head.
The turbine hydraulic efficiency is obviously pooper compared with the pump. When a flow rate is in a very narrow range of
7–10 L/s, the turbine hydraulic efficiency is declined considerably. Fortunately, while the flow rate is above 10 L/s, the turbine
hydraulic efficiency reveals a gentle declination, suggesting that turbine being operated in a higher flow rate is economical. With
the increasing viscosity, the turbine hydraulic efficiency is declined continuously.
The internal variables in pump and turbine modes include the moments of fluid absolute velocity at the impeller inlet and
outlet, the impeller theoretical head, the total hydraulic loss coefficients across the machine and the individual total hydraulic loss
coefficient across the suction pipe/draft tube, impeller and volute. The moments of fluid absolute velocity at the impeller inlet
and outlet and impeller theoretical head are presented in Fig. 3 in terms of flow rate at five viscosities. The impeller theoretical
head is related to the difference in the moment of momentum of fluid across the impeller according to the Euler equation for
turbomachinery, i.e.
⎧
⎪ 1
⎨Hep = (Vu2p u2p − Vu1p u1p ) pump mode
g
(6)
⎪
⎩Het = 1 (Vu1t u1t − Vu2t u2t ) turbine mode
g
Note that the difference in the moment of momentum of fluid, Vu1t u1t − Vu2t u2 , is considered as an internal variable in [29] by
Singh and Nestmann firstly, here we just follow this idea because the impeller theoretical head is equivalent to that difference.
It is shown that the moment of absolute velocity at the impeller inlet in pump mode, rVu1p , is not only tiny but little dependent
on viscosity and flow rate. In turbine mode, the moment of absolute velocity at the impeller outlet, rVu2t , is substantial and largely
dependent on the flow rate but slightly on the viscosity.
The moment of absolute velocity at the impeller outlet in pump mode, rVu2p , depends on flow rate and viscosity considerably,
and its peak values occur at BEPs. Specially, the higher the viscosity is, the larger the rVu2p is.
In turbine mode, the moment of absolute velocity at the impeller inlet, rVu1t , is largely dependent on flow rate and viscosity,
especially at a viscosity lower than 24 cSt. The higher the viscosity, the lower the rVu1t . There is an obvious minimum in the rVu1t
W.-G. Li / Applied Mathematical Modelling 40 (2016) 904–926 911
Fig. 4. Total hydraulic loss coefficient across the whole machine, individual total hydraulic loss coefficient in suction pipe/draft tube, impeller and volute in terms
of flow rate in pump and turbine modes at five viscosities, (a) whole machine, (b) suction pipe/draft tube, (c) impeller, and (d) volute.
curve at a viscosity of 24 cSt or above. Compared with the pump mode, the rVu1t value is smaller than the rVu2p in a wide range of
flow rate, especially for viscosity more than 24 cSt. This suggests the volute is not an efficient vortex generator for more viscous
liquids than water.
The impeller theoretical head in pump mode, Hep , in terms of flow rate and viscosity, resembles the profiles of the moment
of absolute velocity, rVu2p . In turbine mode, because of the effect of the moment of absolute velocity, rVu1t , at the impeller inlet,
the impeller theoretical head, Het , exhibits an obvious influence caused from the viscosity, namely the higher the viscosity, the
lower the theoretical head.
The total hydraulic loss coefficient across the whole machine, individual total hydraulic loss coefficient respectively in the
suction pipe/draft tube, impeller and volute are illustrated in Fig. 4 in pump and turbine modes. Those hydraulic loss coefficients
are defined by the following equations in pump mode
⎧
⎨ξsp = ghsp , ξ = ghip , ξv p = ghv p
ip
u22p u22p u22p (7)
⎩
ξ p = ξsp + ξip + ξv p
And in turbine mode
⎧
⎨ξ = ghst , ξ = ghit , ξ = ghvt
st it vt
u21t u21t u21t (8)
⎩
ξt = ξst + ξit + ξvt
Maybe there are other sorts of definition for hydraulic loss coefficients. Here we just take the impeller characteristic speed to
make the hydraulic losses dimensionless so as to make comparison possible between pump and turbine modes based on the
same scale.
For the total hydraulic loss coefficient across the whole machine, see Fig. 4(a), they reach their minimum at BEP and get the
highest values at a low flow rate in pump mode and at a high flow rate in turbine mode. At the viscosity of 1 cSt, two coefficients
are even larger than those at the other viscosities at a lower flow rate in pump mode and at a higher flow rate in turbine mode.
As the viscosity is equal or more than 24 cSt, the two coefficients augment with the increasing viscosity at a flow rate.
912 W.-G. Li / Applied Mathematical Modelling 40 (2016) 904–926
(d)
(a)
(e)
(b)
(c) (f)
Fig. 5. Wall shear stress on impeller shroud, hub, blades and volute in pump and turbine modes at BEP under 1 cSt, 48 cSt and 120 cSt, (a) 1 cSt, pump, (b) 48 cSt,
pump, (c) 120 cSt, pump, (d) 1 cSt, turbine, (e) 48 cSt, turbine, (f) 120 cSt, turbine, the legend in the left represents wall shear stress in Pa.
Note that the total loss coefficient in turbine mode at BEP or above is always more substantial than in pump mode, suggesting
the turbine is in a poorer hydraulic efficiency as shown in Fig. 2(c).
In terms of the total hydraulic loss coefficients in the suction pipe/draft tube, see Fig. 4(b), the coefficient is less dependent
on the flow rate in pump mode, however, it rises remarkably as the flow rate is beyond 10 L/s in turbine mode. At the viscosity of
1 cSt, in the turbine mode, the coefficient is above the coefficients at a higher viscosity.
The total hydraulic loss coefficients across the impeller in both modes in Fig. 4(c) are quite similar to those across the whole
machine, especially in the pump mode, implying it is the impeller that controls the total hydraulic loss of the machine. In the
turbine mode, the coefficients at both sides of BEP are comparable in magnitude. Unfortunately, the coefficient at 1 cSt viscosity
gets even large at a higher flow rate.
The total hydraulic loss across the volute in pump mode depends on flow rate slightly as shown in Fig. 4(d), in turbine mode,
however, the coefficient steadily rises with the increasing flow rate.
In pump mode, the total hydraulic loss in the suction pipe is so small that it can be ignored; in turbine mode, however, the
loss in the pipe/draft tube is higher than in the volute, so it is not negligible.
The skin friction factor can be important in low Reynolds number flows. Before determining a skin friction factor in pump
and turbine modes, the wall shear stress contours on shroud, hub and blade surfaces are illustrated in Fig. 5 at 1 cSt, 48 cSt and
120 cSt under BEP conditions. Clearly, the wall shear stress level increases with the increasing viscosity in pump and turbine
modes. Further the wall stress in the turbine mode is higher than in the pump.
In both modes, there are some isolated high wall shear stress spots dispersed in the zones where the wall curvatures change
rapidly. In the most areas, however, the stress is uniform. At 0.8BEP and 1.2BEP, the same situation is presented. Therefore, we
can make use of the area averaged wall shear stress to reduce the high stress level contributions and account for the skin fraction
W.-G. Li / Applied Mathematical Modelling 40 (2016) 904–926 913
Fig. 6. Area averaged skin friction factor across suction pipe/daft tube, impeller and volute at 0.8BEP, BEP and 1.2BEP in pump and turbine modes in terms of
Reynolds number, (a) 0.8BEP, (b) BEP, (c) 1.2BEP.
effect in the suction pipe/draft tube, impeller and volute with certain accuracy. Accordingly, an area averaged skin friction factor
can be calculated with the following expressions,
⎧ τsp τip τv p
⎪
⎨ fsp = ρ u2 , fip = ρ u2 , fv p = ρ u2 pump mode
2p 2p 2p
(9)
⎪ τst τit τvt
⎩ fst = 2 , fit = 2 , fvt = 2 turbine mode
ρ u1t ρ u1t ρ u1t
The area averaged skin friction factors across the suction pipe, impeller and volute in pump and turbine modes are illustrated in
Fig. 6 as a function of impeller Reynolds number at 0.8BEP, BEP and 1.2BEP.
Firstly, at the same Reynolds number, a higher flow rate corresponds a larger averaged skin friction factor in each through-
flow component. Secondly, the Reynolds number 5.5151 × 104 (at 24 cSt) appears to be a critical one since the skin friction
factor curve exhibits difference slopes when a Reynolds number is lower or higher than this value, see Table 2, especially for the
friction factor fsp or fst in the suction pipe or draft tube in pump or turbine mode. Because when the Reynolds number is less
than 5.5151 × 104 , the slopes of those curves in Fig. 6 are close to the slope of -1 of a laminar pipe flow, the flow regime in the
pipe in the two modes may be laminar once the liquid viscosity is beyond 24 cSt. However, further experimental conformation
is desirable.
In the pump mode, at 1 cSt viscosity, the skin friction factor in the suction pipe, fsp , is comparable to those in the impeller and
volute. However, with the increasing viscosity or decreasing Reynolds number, even though fsp grows steadily, its value is still so
914 W.-G. Li / Applied Mathematical Modelling 40 (2016) 904–926
Table 2
Slopes of two segments in an area averaged skin friction factor curve divided by the critical Reynolds
number 5.5151 × 104 .
small that can be neglected against those in the impeller and volute. The averaged skin friction factor in the impeller is slightly
larger than in the volute.
In the turbine mode, the averaged skin friction factor in the draft tube, fst , is above the skin friction factor in the volute, fvt , at
0.8BEP and over the skin friction factor in the impeller, fit , at both BEP and 1.2BEP at 1 cSt viscosity. With the increasing viscosity
or decreasing Reynolds number, fst rises and approaches fit and fvt , suggesting fst no longer can be ignored.
In the pump mode, the averaged skin friction factor across the impeller is comparable to that in the volute, but in the turbine
mode, the skin friction factor in the impeller is 1.3–2.2 times higher than in the volute.
The pressure coefficients on the blade suction and pressure sides in the mid-span plane of Blade 3 in Fig. 1(a) and (b) are
illustrated in Fig. 7 at 0.8BEP, BEP and 1.2BEP in pump and turbine modes for 1 cSt, 48 cSt and 120 cSt. In both modes, the
pressure coefficients are defined by the following equations
⎧ p − p1pm
⎪
⎨Cp = 1 ρ u2 pump mode
2 2p
(10)
⎪ p − p2tm
⎩Ct = 1 2 turbine mode
2
ρ u1t
Note that Blade 3 is farthest to the volute tongue and its effect on the pressure coefficient profiles is the least hopefully.
In pump mode, the pressure coefficient on both sides of the blade depends on the viscosity and working condition, especially
on the suction side. The area enclosed by the coefficient is shrunk with the increasing viscosity, causing a declined lift/loading on
the blade with the increase of viscosity. Also, the maximum coefficient reduces with the increasing viscosity. Nearly half length
of the suction side is covered by a negative coefficient starting from the leading edge, suggesting the impeller may be subject to
a poorer cavitation behaviour.
In turbine mode, the pressure coefficients move up considerably with the increasing viscosity, however, the area enclosed by
them does not appear to change with viscosity. This effect indicates that the lift of the blade remains unchanged approximately,
and as if the flow resistance or the pressure difference across the impeller blade row is raised only, particularly at 0.8BEP and
BEP. Further, the blade suction side is subject to a very short length with negative pressure coefficient, suggesting the impeller
will be in a better cavitation performance compared with the pump mode.
Currently, the effects of liquid on centrifugal pump performance are usually handled by three correction factors, namely flow
rate, head and hydraulic efficiency correction factors at 0.8BEP, BEP and 1.2BEP in engineering. By using them, one can obtain the
pump performance when pumping a liquid other than water based on a known performance under water.
For a pump as turbine, there have not been such correction factors so far. To make performance conversion between water
and viscous oil possible like a centrifugal pump, three correction factors in turbine mode are defined as follows at BEP,
⎧
⎪ Qt Ht η Het
⎨KQt = , KHt = , Kηt = ht , Ket = turbine mode
Qtw Htw ηhtw Hetw
(11)
⎪ η
⎩KQ p = Q p , KH p = Hp , Kη p = hp , Kep = Hep pump mode
Q pw Hpw ηhpw Hepw
W.-G. Li / Applied Mathematical Modelling 40 (2016) 904–926 915
Fig. 7. Pressure coefficient on blade surfaces of blade 3 shown in Fig. 1(a) and (b) at 0.8BEP, BEP and 1.2BEP for 1 cSt, 48 cSt and 120 cSt in pump and turbine
modes, (a) 0.8BEP and 1 cSt in pump mode, (b) 0.8BEP and 1 cSt in turbine mode, (c) BEP and 48 cSt in pump mode, (d) BEP and 48 cSt in turbine mode, (e) 1.2BEP
and 120 cSt in pump mode, (f) 1.2BEP and 120 cSt in turbine mode.
916 W.-G. Li / Applied Mathematical Modelling 40 (2016) 904–926
1.5 1.5
1.3 1.3
KHt KHt
1.2 1.2
KQp,KHp,Kηp ,KQt,KHt,K ηt
KQp,KHp,Kηp ,KQt,KHt,Kηt
1.1 KQt 1.1 KQt
1 1
KHp K
KQp KQp Hp
0.9 0.9
Kηt Kηt
0.8 0.8
K ηp Kηp
0.7 0.7
0.6 0.6
4 5 6 4 5 6
10 10 10 10 10 10
Re Re
1.5 1.2
1.3 1.2BEP
1.1
BEP
KQp,KHp,Kηp ,KQt,KHt,Kηt
1.2 KHt
0.8BEP Kep
1.1 KQt
Kep, Ket
1 0.8BEP
1
KHp
KQp
0.9
BEP
0.8 Kηp 0.9 Ket
K ηt 1.2BEP
0.7
0.6
4 5 6 0.8
10 10 10 4 5 6
Re 10 10 10
Re
Fig. 8. Flow rate, head, hydraulic efficiency and impeller theoretical head correction factors at 0.8BEP, BEP 1.2BEP are in terms of impeller Reynolds number in
pump and turbine modes, line-curve fitting, symbol-CFD prediction, the solid lines are for the turbine mode, the dashed lines for the pump mode, (a) 0.8BEP,
(b) BEP, (c) 1.2BEP, and (d) theoretical head correction factor.
where the numerator represents a performance parameter of viscous oil at its own BEP, and the denominator stands for a coun-
terpart parameter of water at its own BEP as well. For comparison, the correction factors in pump mode as well as the impeller
theoretical head correction factors in both mode are involved in Eq. (11), too.
Similarly, the correction factors can be defined at 0.8BEP and 1.2BEP, however, the two flow rate correction factors are identical
to that at BEP, and thus there are head and hydraulic efficiency correction factors only under each of these operating conditions.
Those correction factors in pump and turbine modes are illustrated in Fig. 8. For use in engineering, the factors have been
correlated to impeller Reynolds number by using the following empirical expression at 0.8BEP, BEP and 1.2BEP
Re m
KQ p , KH p , Kη p , Kep , KQt , KHt , Kηt , Ket = A + B (12)
Rew
where the curve fitting coefficient, A and B, and power, m, for those correction factors are listed in Table 3. Actually, Eq. (12) was
proposed in [35,36] to correlate to the total efficiency of inflow gas turbine to its impeller Reynolds number.
The flow rate, head, hydraulic efficiency and theoretical head correction factors-Reynolds number curves are shown in Fig. 8.
When the pump handles viscous oils or more viscous liquid other than water, its BEP is moved to a lower flow rate, its head is
W.-G. Li / Applied Mathematical Modelling 40 (2016) 904–926 917
Table 3
Curve fitting coefficients and power for performance correction factors.
Mode Flow rate Correction factor Coefficients, power and correlation coefficients
A B m R2
degraded and its hydraulic efficiency is reduced. Accordingly, the three correction factors are all less than one. In the poorest
case, the flow rate and head can be reduced by 20%, and the hydraulic efficiency can be degraded by 30%.
When water or viscous oil drives the pump in the reverse direction as a turbine, its BEP is relocated to a higher flow rate, its
head is raised, but its hydraulic efficiency is declined. Hence the flow rate and head correction factors are more than one, but the
hydraulic efficiency correction factor is always smaller than one. In the worst case, the flow rate can be increased by 10%, the head
can be higher by 40%, and the hydraulic efficiency can be lowered by 40%. This implies the effects of viscosity on performance in
the turbine mode are more significant than in the pump mode.
The curve fitting coefficients in Table 3 do not exhibit any remarked differences at 0.8BEP, BEP and 1.2BEP in both modes.
However, the impeller theoretical correction factors shown in Fig. 8(d) change quite significantly from 0.8BEP to 1.2BEP, especially
in turbine mode. For example, at 1.2BEP the theoretical head correction factor can be increased by 18% in pump mode and
decreased by 15% in turbine mode.
4. Discussion
The mass-weighted average absolute velocity, V0 , circumferential component, V0u , and flow angle, α0 , at the volute outlet are
illustrated in Fig. 9 in terms of flow rate. Note that the absolute velocity and its circumferential component share the same profile.
Further, the velocity magnitude steadily increases with flow rate, especially at 1 cSt viscosity, a higher viscosity can result in a
lower velocity. The flow angle decreases with flow rate and viscosity.
It is interesting to notice that the absolute velocity magnitude at 24 cSt exhibits a strange behaviour, namely the velocity is
close to the velocity at the viscosity above 48 cSt at BEP (around 12.01 L/s) and subsequently approaches to the velocity at 1 cSt
at a high flow rate of 18.68 L/s.
To disclose the underlying mechanism of this effect, the absolute velocity vectors in the mid-span plane of the volute in
turbine mode are demonstrated in Fig. 10 at 12.01 L/s and 18.68 L/s for 1 cSt, 24 cSt and 48 cSt. At 1 cSt and 48 cSt, the vector
patterns at the two flow rates remain unchanged essentially. At 24 cSt, however, the pattern at 12.01 L/s flow rate is like that at
48 cSt, while the pattern at 18.68 L/s resembles to that at 1 cSt.
At 1 cSt, the fluid passes by the tongue smoothly without any noted turn and goes into the volute from the nozzle. At 48 cSt,
however, it experiences a sharp turn when passing by the tongue, causing a portion of fluid directly flows into the impeller
channel nearby. As a result of this, the fluid is subject to a higher velocity magnitude and a larger flow angle at 1 cSt than at
48 cSt.
Because the circumferential component, V0u has a principal contribution to the moment of fluid absolute velocity, rVu1t , and
impeller theoretical head, Het , they exhibit a very similar pattern to V0u against flow rate at various viscosities.
918 W.-G. Li / Applied Mathematical Modelling 40 (2016) 904–926
Fig. 9. Magnitude, circumferential component and flow angle of circumferentially averaged absolute velocity at volute outlet in turbine mode at various flow
rate and viscosities, (a) magnitude, (b) circumferential component, (c) flow angle.
Because the sharp turn of fluid around the tongue is a local phenomenon only, unlikely generates a considerable impact on
the hydraulic loss across the volute. The hydraulic loss coefficient steadily rises with increasing viscosity at a flow rate as shown
in Fig. 4(d) is mainly due to the increase in the skin friction factor with reducing Reynolds number as shown in Fig. 6.
In Fig. 2, at 1 cSt viscosity, the turbine hydraulic efficiency declines very quickly with flow rate above the BEP. Further, from
Fig. 4(a) and (c), we have learnt that the total hydraulic loss coefficients across the machine, the total loss coefficients across the
draft tube and impeller rise sharply with increasing flow rate, especially for the impeller. Based on the averaged skin friction
factors shown in Fig. 6, the factor in the impeller at 1 cSt viscosity is quite normal compared with those at the other viscosities.
Thus, it is presumably that the incidence loss and flow separation loss in the impeller may be responsible for that effect.
Therefore it is necessary to clarify the incidence loss at the impeller inlet in turbine mode at various viscosities and flow rates.
The volute outlet is the impeller inlet in turbine mode. We can calculate the mass-weighted average relative velocity of fluid,
W1t , and relative flow angle, β1t , based on absolute velocity, V0 , and absolute flow angle, α0 , as well as the impeller inlet tip speed,
u1t . The relative velocity, W1t , and angle of attack, β1t , incidence loss, hin , and incidence loss coefficient, ξin , are illustrated in
Fig. 11. The incidence losses at various viscosities and flow rates are estimated by using Todd and Futral empirical formula
published in 1969 for gas turbines shown as flows in [37]
W1t2
9.81sin (β1t − φ)
2
hin = (13)
2g
W.-G. Li / Applied Mathematical Modelling 40 (2016) 904–926 919
(a) (b)
(d)
(c)
(e) (f)
Fig. 10. Absolute velocity vectors in mid-span plane of volute at flow rate 12.01 L/s and 18.68 L/s for 1 cSt, 24 cSt and 48 cSt in turbine mode, (a) 1 cSt and
12.01 L/s, (b) 1 cSt and 18.68 L/s, (c) 24 cSt and 12.01 L/s, (d) 24 cSt and 18.68 L/s, (e) 48 cSt and 12.01 L/s, (f) 48 cSt and 18.68 L/s.
where φ is the flow deviation angle from a blade at the leading edge due to the slip effect in turbine mode is estimated by
using the Stodola slip factor formula; note that φ < 0. It is confirmed that incidence loss predicted with Eq. (13) must be
increased by one fold to match the experimental data in [37] for a gas turbine with blunt leading edge. Here, the incidence loss
estimated with Eq. (13) has been amplified by 1.5 times in Fig. 11(c). It is seen in Fig. 11(c) that the incidence loss at 1 cSt is the
highest compared with those at the other viscosities.
The incidence loss coefficient, ξin (=ghin /u21t ) plus the minimum total hydraulic loss coefficient at each viscosity, ξit,min ,
is plotted in Fig. 11(d). For comparison, the total loss coefficient, ξit , is involved in terms of flow rate in that figure as well.
Even though the incidence loss coefficient plus the minimum total hydraulic loss coefficient curve is unable to match the total
hydraulic loss coefficient exactly at each viscosity, it is confirmed that the incidence loss at 1 cSt is the most significant than
those at the other viscosities. This evidence is supportive to that the total loss coefficients across the impeller exhibits a rapid
increase with flow rate at 1 cSt as shown in Fig. 4(c).
At the moment, we are unable to estimate flow separation loss in the impeller quantitatively. However, we can have a
qualitative assessment of the loss by observing the location and size of separation zone alternatively. In doing so, the relative
flow vectors in the mid-span plane of impeller are illustrated in Fig. 12 under the part-load (7.34 L/s), BEP (12.01 L/s) and
over-load (18.68 L/s) conditions at 1 cSt, 24 cSt and 120 cSt viscosities.
When the flow rate is below 12 L/s, the angle of attack is negative, and the liquid impinges the blade suction side directly,
causing a resistant effect to impeller rotation to some degree and a flow separation from the blade pressure side in the rear por-
tion of blade. Consequently, the turbine shows a smaller shaft power and a poorer hydraulic efficiency. As the flow rate is above
12 L/s, the stagnation point of flow is on the pressure side of blade, and the angle of attack is positive, the fluid flow is favour of
920 W.-G. Li / Applied Mathematical Modelling 40 (2016) 904–926
Fig. 11. Averaged relative velocity, angle of attack, incidence loss and its loss coefficient at blade leading edge in turbine mode in terms of flow rate at five
viscosities, (a) relative velocity, (b) angle of attack, (c) incidence loss, (d) incidence loss coefficient, and (e) definition of angle of attack.
impeller rotation, and the turbine can develop a larger shaft power. Unfortunately, there is a reverse flow zone near the pressure
side of blade and the blade leading edge, so that the turbine performance is not perfect. As the flow rate is above 12.01 L/s, even
though the liquid can flow into the impeller smoothly, the fluid still separates from the pressure side of blade in the mid-portion
of blade. Note that those vector plots do reveal that there is a flow separation at any working conditions in the turbine.
Comparing the relative velocity vector plots of water and viscous oil at the same flow rate, it is identified that the size of
separate zone of viscous oil seems smaller than water. The reason for this is that the angle of attack of viscous oil is less than
that of water at the same flow rate, as shown in Fig. 11(b). Therefore, the intensive vortex in the impeller in the turbine mode
should be responsible for the sharp ascending in the total hydraulic loss coefficient at 1 cSt viscosity.
W.-G. Li / Applied Mathematical Modelling 40 (2016) 904–926 921
Fig. 12. Relative velocity vector in mid-span of impeller in turbine mode at low (7.34 L/s), BEP (12.01 L/s) and high (18.68 L/s) flow rates for 1 cSt, 24 cSt and
120 cSt, (a) 1 cSt and 7.34 L/s, (b) 1 cSt and 12.01 L/s, (c) 1 cSt and 18.68 L/s, (d) 24 cSt and 7.34 L/s, (e) 24 cSt and 12.01 L/s, (f) 24 cSt and 18.68 L/s, (g) 120 cSt and
7.34 L/s, (h) 120 cSt and 12.01 L/s, (i) 120 cSt and 18.68L/s.
Relying on the observations above, we can conclude that the flow separation loss in the impeller is larger in the turbine mode
than in the pump mode. Also, the separation loss at a higher viscosity is smaller than in a lower viscosity in the turbine mode.
4.3. Flow condition and deviation angle at impeller outlet in turbine mode
The mass-weighted average circumferential and axial velocities, helix angle and dimensionless pressure profiles along radius
in the impeller outlet/draft tube inlet are illustrated in Fig. 13 at five liquid viscosities and under 0.8BEP, BEP and 1.2BEP condi-
tions. Here the helix angle is defined the angle between the velocity vector and the circumferential direction of the tube which
is in the reverse direction of impeller rotation. The sign ‘-’ for the circumferential velocity stands for the direction is opposite to
the impeller rotational direction. The sign ‘-’ for the axial velocity means the liquid flows into the draft tube, otherwise into the
impeller. Since the radial velocity is as small as 1/20 order of the circumferential velocity, it is no longer presented. The static
pressure has been normalized by means of the averaged static pressure across the draft tube inlet.
922 W.-G. Li / Applied Mathematical Modelling 40 (2016) 904–926
Fig. 13. Mass-weighted average circumferential, axial, helix angle and dimensionless pressure profiles along radius in impeller outlet/draft tube inlet at 0.8BEP,
BEP and 1.2BEP as well as five viscosities, (a) circumferential velocity, (b) axial velocity, (c) helix angle, (d) dimensionless pressure.
It is seen that the circumferential and axial velocities, helix angle and pressure profiles largely depend upon working condi-
tion, and change rapidly across blade span. Particularly, the higher the flow rate, the bigger the peak magnitude of the velocities,
the sharper the velocity profiles and the more uneven the pressure profiles. The velocity and helix angle profiles at 1 cSt are dis-
tinctly different from those at the viscosity of 24 cSt or more, otherwise the velocity profiles are similar across various viscosities.
Since the helix angle varies across the blade span, the flow deviation angle must change as well. Due to boundary layer effect,
the deviation angle near shroud and hub may not be more important than that on the mid-span surface. Consequently, the flow
deviation angle in the impeller outlet on the mid-span surface is shown in Fig. 14 as a function of liquid viscosity at 0.8BEP, BEP
and 1.2BEP. The deviation angle decreases with the reducing flow rate and but gets large with the increasing viscosity. However,
the change in deviation angle against the viscosity is less than against the flow rate. At 0.8BEP, the axial velocity profile is quite
different from those at BEP and 1.2BEP, giving rise to a quite small deviation angle.
4.4. Reason for large hydraulic loss coefficient in draft tube in turbine mode
In Fig. 4(b), the total hydraulic loss coefficient across the draft tube shows a rapid rise with in the increasing flow rate and
becomes comparable with the total loss coefficients in the impeller and volute, indicating the loss coefficient in the tube must
be considered in turbine mode.
W.-G. Li / Applied Mathematical Modelling 40 (2016) 904–926 923
Fig. 14. Flow deviation angle in impeller outlet in terms of liquid viscosity under 0.8BEP, BEP and 1.2BEP conditions on mid-span surface.
Fluid rotation
Fig. 15. Fluid pathlines in turbine mode at 0.8BEP, BEP and 1.2BEP under 1 cSt and 120 cSt, (a) 1 cSt and 0.8BEP, (b) 1 cSt and BEP, (c) 1 cSt and 1.2BEP, (d) 120 cSt
and 0.8BEP, (e) 120 cSt and BEP, (f) 120 cSt and 1.2BEP.
To uncover the reason for such a rapid growth of the coefficient, the fluid pathlines through the machine are illustrated in
Fig. 15 at 0.8BEP, BEP and 1.2BEP under 1 cSt and 120 cSt. The pathlines exhibit a helix pattern in the draft tube in all the case, and
the rotational direction of fluid is opposite to that of the impeller. The helix angle is decreased with increasing flow rate, showing
the enlarged circumferential velocity in magnitude. The generated circumferential component of the velocity can produce an
extra skin friction loss to augment the total hydraulic loss in the draft tube as the swirling flow in the suction pipe of a mixed-
flow pump measured in [38] does.
The total hydraulic loss coefficient and total skin friction factor across draft tube/suction pipe in terms of flow rate at five
viscosities in turbine and pump modes are indicated in Fig. 16. It is shown that the total loss coefficient is 2.1–2.5 times the skin
friction factor in turbine mode, suggesting there are other sorts of hydraulic loss in the draft tube except the total skin friction
factor.
924 W.-G. Li / Applied Mathematical Modelling 40 (2016) 904–926
Fig. 16. Total hydraulic loss coefficient and skin friction factor across draft tube/suction pipe in terms of flow rate at five viscosities in turbine and pump modes,
(a) turbine mode, (b) pump mode.
Fig. 17. Fluid pathlines in pump mode at 0.8BEP, BEP and 1.2BEP under 1 cSt and 120 cSt, (a) 1 cSt and 0.8BEP, (b) 1 cSt and BEP, (c) 1 cSt and 1.2BEP, (d) 120 cSt
and 0.8BEP, (e) 120 cSt and BEP, (f) 120 cSt and 1.2BEP.
However, the total hydraulic loss coefficient in the tube is lightly smaller (the maximum error is −7%) than the total skin
friction factor in pump mode, implying there are no other kinds of hydraulic loss in the suction pipe. The reason for this is that
there is no swirling flow in the suction pipe in this low specific speed centrifugal pump based on the pathline patterns shown in
Fig. 17 at 0.8BEP, BEP and 1.2BEP for 1 cSt and 120 cSt. As indicted in [39], a low specific speed centrifugal pump impeller is not
prone to induce a swirling flow in its suction pipe.
In Fig. 4(d), there are negative total hydraulic loss coefficients in the volute in turbine mode at low flow rate indicated by a
triangle in dashed-line. This means a portion of fluid comes back from the impeller and has gained energy from there rather than
discharges the energy to it, or in other words, there is a ‘pumping effect’ at low flow rate.
At a low flow rate, the blade leading edge is subject to a negative angle of attack of as big as −14° at 1 cSt and −12° at 120 cSt,
see Fig. 11(b); in consequence, the fluid attacks the blade suction (concave) side near the leading edge initially and then it is
deflected outwards by the edge, as shown in Fig. 12(a), (d) and (g), especially at 1 cSt.
W.-G. Li / Applied Mathematical Modelling 40 (2016) 904–926 925
(b)
(a)
Fig. 18. Velocity vectors in mid-span plane of volute in turbine mode at part-load point 7.34 L/s under 1 cSt and 120 cSt viscosities, (a) 1 cSt, (b) 120 cSt.
To confirm the deflected fluid does go back to the volute, the velocity vectors in the mid-span plane of volute are displayed
in Fig. 18 in turbine mode at part-load point of 7.34 L/s under 1 cSt and 120 cSt viscosities. It is clear that there are two streams
are issued from two blade suction sides and flow into the volute, then propagate circumferentially downstream until re-entering
the impeller. The fluid brings some energy of the impeller into the volute and raises the energy level there, causing a negative
total hydraulic loss coefficient in the volute. Because this effect is much stronger at 1 cSt than 120 cSt, a negative coefficient has
to persist until a flow rate as high as 10 L/s at 1 cSt. With the increasing viscosity, the flow rate at which the negative coefficient
disappears is reduced.
Note that a negative total hydraulic loss coefficient in the volute in turbine mode at part-loads is characterized by large flow
separation and reverse flows as well as negative angle of attack, see Figs. 11(b) and 12. However, after BEP and into over-load
regions (14–16 L/s) in Fig. 11(b), the angle of attack tends to zero and the flow is perfectly in line with blade profile, consequently
flow separation disappear in Fig. 12. Thus the negative total hydraulic loss coefficient is a very natural phenomenon in turbine
mode at part-loads.
Note that the flow in the two side chambers between the impeller and the casing, unsteady flow effect and radial and axial
thrusts are not taken into account in the simulations. In future, those factors should be involved in simulations, an experimental
validation should be conducted, and methods for designing efficient impeller of pump as turbine are highly desirable. The effect
of wet wall roughness of hydraulic components on the turbine performance and flow pattern at various viscosities is worthy
being attempted.
5. Conclusions
The steady flows of liquid with a variety of viscosities in an experimental centrifugal pump as turbine are studied by using
CFD tool. Effects of liquid viscosity on flow characteristics and hydraulic performance of the turbine are clarified in detail. It is
shown that the flow rate, total head and hydraulic efficiency of the turbine rise with increasing viscosity at BEP, and the output
shaft power is mainly affected by the density of liquid rather than the viscosity. Liquid viscosity has even more adverse effect
on turbine performance than on pump. The hydraulic losses of the turbine mainly occur in the impeller and draft tube. The
flow separation in the impeller and incidence loss are responsible for the turbine having a lower hydraulic efficiency than the
pump. The external performance parameters and internal variables, impeller theoretical head and correction factors in both
pump and turbine modes are related to impeller Reynolds number very well by using a three-parameter-power function. The
increased wall shear stress is responsible for the degraded turbine performance at a higher viscosity. The draft tube is subject to
a strong swirling flow at any flow rates and viscosities in turbine mode. The forthcoming work will be the turbine performance
measurement, impeller and volute geometrical effects on the turbine performance and flow pattern at various viscosities.
926 W.-G. Li / Applied Mathematical Modelling 40 (2016) 904–926
References
[1] N. Koyama, Utilization of centrifugal pumps used as hydraulic turbines, J. Japan Soc. Mech. Eng. 74 (635) (1971) 1584–1587.
[2] B. Orchard, Pumps as turbines in water industry, World Pumps 8 (2009) 22–23.
[3] H. Ramos, A. Borga, Pumps as turbines: an unconventional solution to energy production, Urban Water 1 (1999) 261–263.
[4] M. Arriaga, Pump as turbine – a pico-hydro alternative in Lao People’s Democratic Republic, Renew. Energy 35 (2010) 1109–1115.
[5] T. Agarwal, Review of pump as turbine (PAT) for micro-hydropower, Int. J. Emerg. Technol. Adv. Eng. 2 (11) (2012) 163–169.
[6] H. Nautiyal, V. Anoop Kumar, Reverse running pumps analytical, experimental and computational study: a review, Renew. Sustain. Energy Rev. 14 (2010)
2059–2067.
[7] S.V. Jain, R.N. Patel, Investigations on pump running in turbine mode: a review of the state-of-the-art, Renew. Sustain. Energy Rev. 30 (2014) 841–868.
[8] R. Singh, S.V. Cabibbo, Hydraulic turbine energy recovery-R. O. system, Desalination 32 (1980) 281–296.
[9] W.A. Raja, R.W. Piazza, Reverse running centrifugal pumps as hydraulic power recovery turbines for seawater reverse osmosis systems, Desalination 38
(1980) 123–134.
[10] S. Gopalakrishnan, Power recovery turbines for the process industry, in: Proceedings of the 3rd International Pump User Symposium, Houston, USA, 1986.
[11] A.A. Williams, The turbine performance of centrifugal pumps: a comparison of prediction methods, Proc. Inst. Mech. Eng. A J. Power Energy 208 (1984)
59–66.
[12] S. Derakhshan, A. Nourbakhsh, Experimental study of characteristic curves of centrifugal pumps working as turbines in different specific speeds, Exp. Therm.
Fluid Sci. 32 (2008) 800–807.
[13] S. Derakhshan, A. Nourbakhsh, Theoretical, numerical and experimental investigation of centrifugal pumps in reverse operation, Exp. Therm. Fluid Sci. 32
(2008) 1620–1627.
[14] P. Singh, F. Nestmann, An optimization routine on a prediction and selection model for the turbine operation of centrifugal pumps, Exp. Therm. Fluid Sci. 34
(2010) 152–164.
[15] H. Nautiyal, K.A. Varun, S. Yadav, Experimental investigation of centrifugal pump working as turbine for small hydropower systems, Energy Sci. Technol. 1
(1) (2011) 79–86.
[16] J.C. Pascoa, F.J. Silva, J.S. Pinheiro, D.J. Martins, A new approach for predicting PAT-pumping operating point from direct pumping mode characteristics, J. Sci.
Ind. Res. 71 (2012) 144–148.
[17] N. Raman, I. Hussein, K. Palanisamy, B. Foo, An experimental investigation of pump as turbine for micro hydro application, in: IOP Conference Series: Earth
and Environmental Science, 16, 2013, pp. 1–4.
[18] H. Shinhama, J. Fukitomi, Y. Nakase, Y.T. Chen, T. Kuwauchi, S. Miyauchi, Study on reverse running pump turbine, Trans. JSME Ser. B 65 (638) (1999) 3399–
3406.
[19] Y. Nakatake, Y. Kuma, A. Miyao, T. Kurokawa, Study on performance and improvement of efficiency for reverse running pump turbine, Memoirs Kurume
Natl. Coll. Technol. 17 (1) (2001) 1–6.
[20] J. Fernandez, E. Blanco, J. Parrondo, M.T. Stickland, T.J. Scanlon, Performance of a centrifugal pump running in inverse mode, Proc. Inst. Mech. Eng. A J. Power
Energy 218 (2004) 265–271.
[21] S. Rawal, J.T. Kshirsagar, Numerical simulation on a pump operating in a turbine mode, in: Proceedings of the 23rd International Pump User Symposium,
Houston, USA, 2007.
[22] J. Gonzalez, J. Fernandez, K.M. Arguelles-Diaz, C. Santolaria, Flow analysis for a double suction centrifugal machine in the pump and turbine operation
modes, Int. J. Numer. Methods Fluids 61 (2009) 220–236.
[23] S.S. Yang, S. Derakhshan, F.Y. Kong, Theoretical, numerical and experimental prediction of pump as turbine performance, Renew. Energy 48 (2012) 507–513.
[24] H. Nautiyal, Varun, A. Kumar, CFD analysis on pumps working as turbines, 2010, pp. 35–37.
[25] C.S. Morros, J.M.F. Oro, K.M.A. Diaz, Numerical modelling and flow analysis of a centrifugal pump running as a turbine: unsteady flow structures and its
effects on the global performance, Int. J. Numer. Methods Fluids 65 (2011) 542–562.
[26] R. Barrio, J. Fernandez, E. Blanco, J. Parrondo, A. Marcos, Performance characteristics and internal flow patterns in a reverse-running pump–turbine, Proc.
Inst. Mech. Eng. C J. Mech. Eng. Sci. 226 (2011) 695–708.
[27] O. Fecarotta, A. Carravetta, H.M. Ramos, CFD and comparisons for a pump as turbine: mesh reliability and performance concerns, Int. J. Energy Environ. 2
(1) (2011) 39–48.
[28] A. Couzinet, L. Gros, D. Pierrat, Characteristics of centrifugal pumps working in direct or reverse mode: focus on the unsteady radial thrust, Int. J. Rotating
Mach. (2013) Article ID 279049, 11 pp.
[29] P. Singh, F. Nestmann, Internal hydraulic analysis of impeller rounding in centrifugal pumps as turbines, Exp. Therm. Fluid Sci. 35 (2011) 121–134.
[30] S.S. Yang, F.Y. Kong, W.M. Jiang, X.Y. Qu, Effects of impeller trimming influencing pump as turbine, Comput. Fluids 67 (2012) 72–78.
[31] V.A. Patel, S.V. Jain, K.H. Motwani, R.N. Patel, Numerical optimization of guide vanes and reducer in pump running in turbine mode, Procedia Eng. 51 (2013)
797–802.
[32] W.G. Li, Mechanism for onset of sudden-rising head effect in centrifugal pump when handling viscous oils, ASME J. Fluids Eng. 136 (2014) 074501-1-10.
[33] Rohlik H.E., Analytical Determination of Radial Inflow Turbine Design Geometry for Maximum Efficiency, NASA TN D-4384, 1968.
[34] Anonymous, FLUENT 6.2 User’s Guide, Vols. 1 and 2, Fluent Inc, Lebanon, NH 03766, USA, 2005.
[35] Holeski D.E. and Futral S.M., Experimental Performance Evaluation of a 6.02-Inch Radial-Inflow Turbine over a Range of Reynolds Number, NASA TN D-3824,
1967.
[36] Nusbaum W.J. and Wasserbausr C.A., Experimental Performance Evaluation of a 4.59-Inch Radial-Inflow Turbine over a Range of Reynolds Number, NASA
TN D-3835, 1967.
[37] S.W.T. Spence, D.W. Artt, An experimental assessment of incidence losses in a radial inflow turbine rotor, Proc. Inst. Mech. Eng. A J. Power Energy 212 (1998)
43–53.
[38] M. Murakami, N. Heya, Swirling flow in suction pipe of centrifugal pumps (2nd Report, Distribution of Shearing Stress), Bull. JSME 9 (34) (1966) 337–344.
[39] M. Murakami, N. Heya, Swirling flow in suction pipe of centrifugal pumps (1st Report, Distribution of Velocity and Energy), Bull. JSME 9 (34) (1966) 328–337.