You are on page 1of 17

International Journal of Heat and Fluid Flow 75 (2019) 44–60

Contents lists available at ScienceDirect

International Journal of Heat and Fluid Flow


journal homepage: www.elsevier.com/locate/ijhff

Effect of working conditions and diffuser setting angle on pressure T


fluctuations within a centrifugal pump
⁎,a,b
Antonio Posa , Antonio Lippolisc
a
CNR-INM, Institute of Marine Engineering, National Research Council of Italy, Via di Vallerano 139, Roma 00128, Italy
b
Department of Mechanical and Aerospace Engineering, The George Washington University, 800 22nd Street N.W., Washington, DC 20052, USA
c
Dipartimento di Meccanica, Matematica e Management, Politecnico di Bari, Viale Japigia 182, Bari 70126, Italy

ARTICLE INFO ABSTRACT

Keywords: Large Eddy Simulations (LES) were carried out on a centrifugal pump in order to investigate the effect on
Large-Eddy simulation pressure fluctuations through rotating and stationary channels by both flow-rate conditions and orientation of
Immersed-boundary method the diffuser blades. Computations were performed for Reynolds numbers, based on the inflow velocity and outer
Centrifugal pumps impeller radius, equal to Re = 1.5 × 105 (design flow-rate) and Re = 6 × 10 4 (off-design flow-rate, equivalent to
Pressure fluctuations
40% of the nominal one). Results show that sensitivity to both working conditions and setting angle of the
Off-design working conditions
diffuser blades is much stronger for the stationary vanes, compared to the rotating channels. At reduced flow-
rates pressure fluctuations are higher, with more energy at frequencies lower than that of passage of the impeller
blades. However, rotating stall phenomena were not observed. A reduction of the geometric inlet angle of the
diffuser vanes produces an increase of the incidence angle on the diffuser blades at the nominal flow-rate and a
decrease at the lower flow-rate. Pressure rise is negatively impacted in the former case, while the effect is
beneficial in the latter one. In contrast, pressure fluctuations decrease for both flow-rates, mainly due to the
wider gap between trailing edge of impeller blades and leading edge of diffuser blades. With the modified
diffuser geometry, at the design flow-rate the role of the low frequencies keeps negligible and at the reduced
flow-rate they become also less energetic, thanks to improved diffuser inflow and weaker separation phenomena.

1. Introduction the trailing edges. In general, pressure fluctuations were verified of the
same order of magnitude as the pressure rise through the pump, al-
Pressure fluctuations in turbopumps are a major concern, especially though increasing the number of diffuser blades allowed reducing the
at off-design working conditions, due to the structural stress they pro- unsteady behavior of the flow. Yao et al. (2011) performed pressure
duce on the pump and other elements of the system where the pump measurements in the volute casing of a double-suction centrifugal
operates, causing also vibrations, noise and fatigue. Therefore several pump. They found that a flow-rate increase to only 107% of the design
scholars have investigated the unsteady pressure signals generated value caused an increase of the amplitude of pressure fluctuations to
within centrifugal and mixed-flow pumps, performing parametric stu- 254% of the value at nominal conditions. A similar trend was verified
dies and suggesting design solutions aimed at decreasing the unsteady also at reduced flow-rates. A pump-turbine model was studied experi-
features of the pressure field. Experimental studies have provided im- mentally at design and lower flow-rates conditions by
portant information on the influence of both working conditions and Yang et al. (2015). The spectral analysis of pressure signals, coupled
pump geometry on the unsteady nature of the pressure field (Arndt with flow visualizations, allowed the Authors to identify an unsteady
et al., 1990; Sinha et al., 2001; Parrondo-Gayo et al., 2002; Yao et al., phenomenon at a sub-frequency component, associated to back-flow in
2011; Yang et al., 2015; Zhang et al., 2015b; Fu et al., 2018). For in- the return vane channel for both full and part-load conditions. An ad-
stance, Arndt et al. (1990) carried out measurements in both stationary ditional lower frequency perturbation developed at low flow-rates, with
and rotating channels of a centrifugal pump. In the vaned diffuser they rotating stall affecting the diffuser vanes.
found higher pressure pulsations on the suction side and increased Measurements in centrifugal pumps are obviously very challenging
unsteadiness for smaller gaps between impeller and diffuser vanes. and expensive, especially due to the presence of rotating parts.
Measurements in the impeller vanes revealed the largest fluctuations at Therefore, numerical tools are often adopted to gain additional insight,


Corresponding author.
E-mail addresses: antonio.posa@insean.cnr.it, aposa@gwu.edu (A. Posa), antonio.lippolis@poliba.it (A. Lippolis).

https://doi.org/10.1016/j.ijheatfluidflow.2018.11.011
Received 15 July 2018; Received in revised form 24 October 2018; Accepted 24 November 2018
Available online 30 November 2018
0142-727X/ © 2018 Elsevier Inc. All rights reserved.
A. Posa, A. Lippolis International Journal of Heat and Fluid Flow 75 (2019) 44–60

thanks to the wealth of information available from the full three-di- computations, performed using an in-house solver with no numerical
mensional flow fields. Several examples can be found in the literature dissipation and coupled with a more advanced eddy-viscosity model,
(Feng et al., 2007; Spence and Amaral-Teixeira, 2008; 2009; Barrio compared to the classical Smagorinsky model (Smagorinsky, 1963). We
et al., 2010; Gao et al., 2014; 2016; Fu et al., 2015; Zhang et al., 2015a). adopted a computational grid composed of almost 500 million nodes, in
For instance, Feng et al. (2007) computed via unsteady RANS (Rey- order to resolve all important energy-carrying structures of the flow. A
nolds-averaged Navier-Stokes) the flow through a vaned-diffuser radial grid refinement study was also carried out, adopting two additional
pump. In the diffuser vanes they observed the smallest pressure fluc- coarse and medium grids, composed of 100 million and 220 million
tuations at design working conditions. In the impeller vanes oscillations nodes, respectively, demonstrating satisfactory independence from grid
of pressure increased substantially at lower flow-rates. Larger radial resolution for both first and second order statistics. We studied the
gaps between impeller and diffuser blades were found again very influence of flow-rate conditions and setting angle of the diffuser blades
beneficial in decreasing the unsteady nature of the flow. Also on averages, root mean squares (RMSs) and spectra of pressure through
Barrio et al. (2010) utilized unsteady RANS computations, validated both impeller and diffuser vanes of a centrifugal pump, considering
against experimental measurements, to analyze the flow through a both time-averaged and phase-averaged statistics. We utilized a com-
volute-type centrifugal pump across a large range of operating condi- putational strategy already validated on the same flow problem in our
tions, between 20% and 160% of the design flow-rate. The pressure earlier studies (Posa et al., 2011; 2015; 2016a; Posa and Lippolis,
coefficient along the volute front side experienced the highest oscilla- 2018), where we reported also a general discussion on the influence of
tions near the volute tongue, due to interaction with the trailing edge of flow-rate and diffuser geometry on flow physics. In the following this
the impeller blades. This phenomenon was reinforced at flow-coeffi- paper presents: the adopted numerical method (Section 2), the pump
cients both higher and lower than the nominal one. RANS, coupled with geometry (Section 3), the analysis of the pressure signals at probes lo-
experiments, was adopted also by Gao et al. (2014), to study a cen- cated within the impeller and diffuser vanes, together with phase-
trifugal pump with stay vanes. They computed the flow in both steady averaged fields, including also a grid refinement study, (Section 4) and
and unsteady conditions. Away from the nominal flow-rate the results conclusions (Section 5).
from steady simulations deviated significantly from the experiments,
while the agreement was largely improved in the unsteady simulations, 2. Numerical method
with a better prediction of the dominant frequency of pressure signals.
The amplitude of pressure fluctuations at the impeller/diffuser interface In this study we adopted an LES technique to simulate the flow
was found roughly an order of magnitude higher than that at the im- within a centrifugal pump, for two different flow-rate conditions and
peller inlet. Zhang et al. (2015a) carried out computations and ex- three setting angles of the diffuser blades. A general discussion on
periments on a centrifugal pump with slope volute. At the nominal present computations was already reported by Posa and
flow-rate the RANS methodology adopted for the computations was Lippolis (2018), while here we will focus our analysis on pressure
able to predict accurately pressure fluctuations at the blade passage fluctuations within rotating and stationary channels. LES is a time-ac-
frequency, comparing well with the measurements by pressure trans- curate methodology, where turbulence modeling is limited to the
ducers, while large deviations were generated at off-design conditions, structures smaller than grid spacing, in contrast with more conventional
for which strong low-frequency phenomena were produced by flow RANS methods, resolving the time-averaged Navier–Stokes equations
separation. For further examples of RANS studies see, for instance, and using turbulence modeling also for the large scales of the flow.
Pei et al. (2012), Shi and Tsukamoto (2001), Si et al. (2013), However, the accurate modeling of the large scales is very challenging,
Si et al. (2014), Liu et al. (2018) and Zheng et al. (2018). since they are problem-dependent and lack of universality. Therefore,
Based on the literature survey discussed above, we verified that RANS techniques are less computationally demanding, compared to
numerical studies on pressure fluctuations within centrifugal pumps are LES, but also less accurate, especially in presence of separation. In the
usually limited to RANS or – in a very small number of cases – to literature survey reported in Section 1 this feature of RANS was proved
Smagorinsky LES computations (Wang and Wang, 2013; Gao et al., problematic away from the design working conditions, especially at
2016), adopting numerical methods that are better suited for the former flow-rates lower than design, for which the flow experiences strong
time-averaged approaches than for eddy-resolving methodologies, re- separation phenomena, enhancing turbulence levels, thus affecting also
quiring optimal conservation properties of the numerical solver. Few pressure signals within rotating and stationary channels.
LES studies using more sophisticated subgrid scales (SGS) models are The non-dimensional filtered Navier–Stokes equations for in-
also available in the literature, as in Kato et al. (2005, 2007), where the compressible flows were resolved, where filtering is implicitly defined
flow through a five-stages centrifugal pump was simulated adopting a by the resolution of the discretization in space and in time:
dynamic Smagorinsky model, extending the capabilities of the standard
u˜ = 0,
Smagorinsky model within the computational tool developed by Kato
u˜ 1
et al. (2003). However, in Kato et al. (2005, 2007) the computational + ˜ ˜) =
·(uu p˜ · + 2u
˜ + f,
t Re (1)
grid was not designed to resolve turbulent boundary layers, discretizing
the overall fluid domain within the machine by 37 million finite ele- In Eq. (1) the symbol ∼ stands for the filtered quantities. There-
ments, with 0.3 million elements for each vane of the pump. The dy- fore, ũ and p̃ are the resolved velocity vector and pressure, respectively.
namic Smagorinsky model was adopted also by Pacot et al. (2016), to The other quantities are the time variable, t, the subgrid stress tensor, τ,
study the rotating stall phenomenon at part-load conditions in a pump- the Reynolds number, Re, and a forcing term, f. The SGS tensor takes
turbine working in pumping mode. Recently also Kye et al. (2017, into account the scales that cannot be resolved by the adopted resolu-
2018) utilized a more advanced SGS model, the dynamic global model tions in space and in time. Here we utilized the Wall-Adapting Local
developed by Lee et al. (2010), to simulate the internal flow of a cen- Eddy-viscosity (WALE) model by Nicoud and Ducros (1999) to compute
trifugal volute pump at both design and off-design working conditions. τ, based on the square of the velocity gradient tensor of the resolved
Both experimental and computational analyses above have de- velocity field. It is an inexpensive model (few percent of the overall
monstrated that the most substantial unsteady phenomena occur at off- computational cost), but more sophisticated than the Smagorinsky
design conditions, because of massive separation, whose simulation via model, adopted in some studies discussed in Section 1. Since the
RANS methods is problematic. This makes LES a more suitable tool to computation of the subgrid stresses is based not only on the deforma-
investigate numerically pressure fluctuations in turbopumps flows, al- tion tensor, but also on the rotation tensor, the WALE model is able to
beit computationally more demanding, which explains the very limited readjust in laminar flows and in the near wall regions without ad hoc
number of studies available in the literature. Here we present LES corrections, which is a significant advantage in the present case, dealing

45
A. Posa, A. Lippolis International Journal of Heat and Fluid Flow 75 (2019) 44–60

with a complex three-dimensional geometry. Such features were dis-


cussed by Nicoud and Ducros (1999) and verified also in the challen-
ging case of the flow over the hull of a submarine by Posa and
Balaras (2018b), where the eddy viscosity was shown to scale as (y+ )3,
being y+ the distance from the surface of the body in wall-units. The
same SGS model was adopted to simulate this pump (Posa et al., 2015;
2016a; Posa and Lippolis, 2018), demonstrating good comparisons with
the reference measurements. We recall that Posa et al. (2011) utilized
instead the classical Smagorinsky model (Smagorinsky, 1963) and the
Filtered Structure Function (FSF) model (Ducros et al., 1996), proving
the superiority of the latter. However, based on the discussion reported
by Ducros et al. (1998), the near wall behavior of the WALE model is
physically more consistent than that of the FSF model, not scaling or
scaling as (y+ )2 , if the stencil adopted by the model is properly modified
in the near wall region, computing the structure function of the velocity
field only across the directions parallel to the wall. Therefore, the WALE
model was selected to carry out the present study. More details about it
are reported by Nicoud and Ducros (1999). The Reynolds number in
Eq. (1) is defined as Re = UL/ , where U and L are the reference ve-
locity and length scales, while ν is the kinematic viscosity of the fluid.
The forcing term, f, is utilized to enforce the boundary conditions re-
presenting the action of the body on the fluid. To model this term we
adopted an Immersed-Boundary (IB) methodology (Mittal and
Iaccarino, 2005; Peskin, 1977; 2002; Verzicco et al., 2000), where the
body is “immersed” within a regular non-conforming grid (see Fig. 1)
and local interpolations define the velocity boundary conditions at the
nodes of the Eulerian grid of interface between fluid and solid regions of
the domain. In the current implementation of the IB method, whose
details can be found in Balaras (2004), Yang and Balaras (2006) and
Vanella et al. (2014), the interface nodes are those outside of the body,
having at least one neighboring node inside the body along any of the
three coordinate directions in space. Linear interpolations along the
direction normal to the surface of the immersed-boundary were utilized
to reconstruct the solution at the interface nodes, based on the no-slip
wall boundary condition on the solid body and the solution within the
Fig. 1. (a) Rotating and stationary channels of the pump within a r slice of
fluid region of the computational domain. The surface of the immersed-
the cylindrical computational grid. For visibility purposes impeller and diffuser
boundary was discretized by triangular elements, using a Stereolitho- blades, as well as casing, were clipped over the plane of the slice and only 1 of
graphic (STL) format. Note that the IB method is especially convenient every 8 grid nodes is shown, along both azimuthal and radial directions. (b)
in the present case, involving elements in relative motion, as demon- Detail at the interface between impeller and diffuser, in the area of the dashed
strated in several scenarios featuring moving bodies (Balaras et al., box in panel (a). Only 1 of every 2 grid nodes is shown.
2015; Barsky et al., 2014). Relaxing the requirement for the Eulerian
grid to conform to the body, no grid deformation or regeneration is
approach avoids numerical dissipation, which is a critical requirement
necessary during the simulation. Several validations of the present
for explicit LES, where dissipation should come only from the SGS
solver in complex flow problems are available in the literature (Posa
tensor, rather than from the numerical scheme. The advancement in
and Balaras, 2016; 2018a; 2018b; Posa et al., 2016b). Here Eq. (1) was
time utilized a fractional-step method (Chorin, 1967; Van Kan, 1986). A
resolved on a cylindrical grid, which allowed us to cluster more Eu-
semi-implicit approach allowed us to relax the resolution constraints in
lerian nodes in the impeller/diffuser region of the turbopump, where
time. The implicit Crank–Nicolson scheme was utilized for the terms of
gradients and resolution requirements are more demanding. This grid
azimuthal derivative at small radii and for those of radial derivative at
was composed of almost 500 million nodes: 718 × 1602 × 402 along
the impeller/diffuser interface. All other terms were approximated by
the radial, azimuthal and axial directions, respectively. A refinement
the explicit three-steps Runge–Kutta scheme. The value for the Cour-
study was also carried out using cylindrical grids composed of
ant–Friedrichs–Lewy (CFL) number, defining the resolution in time,
400 × 1002 × 252 (100 million) and 600 × 1202 × 302 (220 million)
was chosen equal to 1. The solution of the Poisson problem, stemming
nodes, respectively. Hereafter those grids will be referred to as “coarse”
from the continuity condition, was achieved via a Fast Fourier Trans-
and “medium”, in contrast with the “fine” grid discussed above. At the
form (FFT) along the azimuthal direction. Then the resulting penta-
inflow boundary of the computational domain a Dirichlet condition was
diagonal systems of equations in each meridian plane of the cylindrical
enforced, while a convective condition at the outflow (Orlanski, 1976)
grid were resolved using a fast direct solver (Rossi and Toivanen, 1999),
allowed eddies to be transported away from the domain without cor-
adopting a variant of the cyclic reduction algorithm by
rupting the solution. Note that in our earlier studies (Posa et al., 2011;
Swarztrauber (1974).
2015; 2016a; Posa and Lippolis, 2018) the overall methodology was
validated on the present flow problem using both the same fine grid
considered here and coarser grids, via comparison with PIV measure- 3. Geometry of the pump and working conditions
ments by Boccazzi et al. (2009) at both design and off-design working
conditions. The geometry of the turbopump studied experimentally by
To resolve Eq. (1) we utilized an in-house MPI finite-difference Boccazzi et al. (2009) is presented in Fig. 2. The machine is a cen-
solver, developed by Balaras (2004). The discretization in space was trifugal pump: the impeller inflow from the suction channel (A) is axial,
based on second-order central differences on a staggered grid. This whereas its outflow into the diffuser vanes is mainly radial/azimuthal.

46
A. Posa, A. Lippolis International Journal of Heat and Fluid Flow 75 (2019) 44–60

the working conditions of the pump, are:

BEP Q0.5
s = = 1.08,
(gH )0.75 (2)

BEP Q
= = 0.291,
A1 u1 (3)

BEP gH
= = 0.443,
u22 (4)

where Ω is the impeller rotational speed, Q the volumetric flow-rate, g


the acceleration of gravity, H the head generated by the pump, A1 the
area of the impeller inlet, u1 the tangential velocity at the outer radius
of the impeller inlet and u2 the tangential velocity at the trailing edge of
the impeller blades.
Here the nominal flow-rate condition, corresponding to
Qn = 0.025 [m3/s], as well as a reduced flow-rate,
Qr = 40%Qn = 0.010 [m3/s], are studied. In the latter case a reduction of
Fig. 2. Elements of the simulated centrifugal pump: meridian cut of the suction the geometric angle of the diffuser inlet, β3, was verified beneficial,
pipe (A); six impeller blades (B); hub (C); meridian cut of the impeller shroud decreasing the incidence at the leading edge of the diffuser blades, as
(D); seven diffuser blades (E); diffuser wall on the hub side (F); volute and discussed in detail by Posa and Lippolis, 2018. For all simulations (two
casing on the hub side (G). Locations of inflow and outflow are also shown. flow-rates and three geometries) the rotational speed was the same as in
the experiments by Boccazzi et al. (2009): = 55.4 [rad/s]. As dis-
The impeller is composed of six blades (B), hub (C) and shroud (D). The cussed above, the small dimensions of the machine lead to rather small
pump features seven diffuser blades (E). The walls of the diffuser vanes values of the Reynolds number, making the LES/IB approach feasible.
(F) on the shroud and hub sides, respectively, are parallel. The volute The Reynolds number, based on the inflow velocity and the impeller
(G) has a rectangular cross-section of increasing area from its tongue to outer radius, was Re ≈ 1.5 × 105 at the nominal flow-rate and
the discharge pipe. In the present study two additional diffuser geo- Re ≈ 6 × 104 at the reduced one. Fig. 3 shows the characteristic curves
metries were analyzed, besides that considered by of the original design of the centrifugal pump, as reported by
Boccazzi et al. (2009). They were generated by modifying the setting Boccazzi et al. (2009). In Fig. 3 also the values of head coefficient, ψ,
angle of the diffuser blades. Such modification will be discussed more and efficiency, η, from the present computations carried out on the
carefully later, when the orientation of the diffuser blades for the three three utilized grids are provided. The agreement between experiments
computed configurations will be shown. Several additional details on and computations is satisfactory with all resolutions. Note that the
the pump geometry are reported by Boccazzi et al. (2009) and by evolution of ψ as a function of φ is rather flat, which emphasizes the
Posa et al. (2016a). Here we will limit our discussion to the radii of error between measured and computed values of head coefficient,
impeller and diffuser at inlet and outlet, respectively, to point that we keeping actually below 4% in all cases. For instance, at the design flow-
are considering a small machine, with relatively low values of Reynolds rate the relative error was found equal to 2.03% and 2.39% on the
number, making the LES approach computationally affordable. At the fine, medium and coarse grids, respectively. Grid convergence was
impeller inlet R1 = 77 [mm], at the impeller outlet R2 = 112 [mm], at improved at the off-design flow-rate, for which the relative error was
the diffuser inlet R3 = 116.5 [mm] and at the diffuser outlet equal instead to 0.05%, 1.30% and 3.10% on the three adopted grids,
R 4 = 180.5 [mm]. At the best efficiency point (BEP) the specific speed, respectively. Similar values of error were found for the efficiency of the
Ωs, the flow coefficient, φ, and the head coefficient, ψ, characterizing pump.

Fig. 3. Characteristics of the centrifugal pump in Fig. 2. Lines


from the experiments by Boccazzi et al. (2009), symbols from
the present computations on the original geometry. The dotted
line stands for the head coefficient, whereas the dashed line
for the efficiency. Triangles, squares and diamonds showing
numerical results on the fine, medium and coarse grids, re-
spectively.

47
A. Posa, A. Lippolis International Journal of Heat and Fluid Flow 75 (2019) 44–60

Fig. 4. Detail of the diffuser region, showing the pro-


file of the stationary blades for 3 = 18 (a), 3 = 14 (b)
and 3 = 10 (c).

4. Results mid camber location was considered more straightforward and realistic
for an actual machine, although this way the following analysis in-
In this paper results are presented for pressure fluctuations and their volves the overlapping effects of two variables: the incidence over the
dependence on flow-rate conditions and setting angle of the diffuser diffuser blades and the radial distance between the trailing edge of the
blades. Further details are discussed by Posa and Lippolis (2018), where impeller blades and the leading edge of the diffuser blades. Fig. 5 shows
the focus is mainly on the first and second order statistics of the velocity the internal elements of the centrifugal pump, including also the loca-
fields. The sensitivity of the global performance of the pump to working tion of the pressure probes within impeller and diffuser. The view is
conditions and diffuser geometry is also provided there. Validation via from the pump inlet looking downstream. In Fig. 5 the orientation of
comparisons with experiments by Boccazzi et al. (2009), who carried the diffuser blades is the original one, corresponding to 3 = 18 . The
out PIV measurements on the original pump geometry, is reported by position of the pressure probes in Fig. 5 is shown using + symbols, in
Posa et al. (2011, 2015, 2016a). black for the diffuser probes and in white for the impeller ones. The
The analysis of pressure fluctuations presented here was performed diffuser probes are located at about the mid radius of the stationary
on the pressure signals at probes located within the diffuser and im- channels. Their position was chosen closer to the suction (concave)
peller vanes, with a particular focus on the former, where the influence side, facing directly the rotating channels. There pressure fluctuations
of flow-rate and diffuser geometry on pressure statistics was found more are expected to be higher, due to the stronger interaction with the
substantial. Besides ensemble-averages over time at the impeller and impeller wake, compared to that affecting the pressure (convex) side of
diffuser probes, also phase-averaged fields of pressure RMSs are pre- the diffuser channels (for more details see also the discussion reported
sented, computed using a statistical sample synchronized with the im- by Posa et al., 2011). The nomenclature that will be adopted hereafter
peller rotation, to shed more light on the interaction between moving identifies the diffuser probes as “dαsβ”, where α stands for the azi-
and stationary parts. For all simulations discussed below, after statis- muthal location within the diffuser, shown in Fig. 5, and β for the
tically steady conditions were established, ensemble and phase-aver- spanwise location, as percentage of the span, measured from the hub
aged statistics were computed over 10 impeller revolutions. Three dif- side towards the shroud side. Thus, each symbol within the diffuser
fuser geometries were considered, with the original design having a channels in Fig. 5 stands actually for a row of probes distributed along
diffuser inlet angle equal to 3 = 18 . The modified diffuser geometries the span at 7%, 23%, 50%, 77% and 93%. The impeller probes are at
were generated via rotation of the blades around their mid camber 75% of the rotor outer radius, R2, at an axial location corresponding
location, as shown in Fig. 4, leading to values of 3 = 14 and 3 = 10 , roughly to the mid axial extent of the impeller and are distributed over
respectively, as well as to an enlarger gap between rotating and sta- the azimuth of one of the rotating channels, moving together with the
tionary blades. The choice of rotating the diffuser blades around the impeller. In Fig. 5 the impeller blades were clipped over the plane of the
probes, for visibility purposes. The nomenclature utilized for those
probes is “rγ”, from the pressure side ( = 1) towards the suction side
( = 7 ). Also for the impeller blades the pressure and suction sides
correspond to their convex and concave surfaces, respectively. In Fig. 5,
for limitation of space, only the first (r1) and the last (r7) probes are
explicitly indicated.

4.1. Grid refinement study

As discussed above, independence of results from grid resolution


was verified via computations on three different grids, composed of
100, 220 and 500 million nodes. Grid refinement was tested for both
working conditions and all three diffuser setting angles. For the sake of
conciseness, results of these tests will be presented here only for the off-
design condition, since grid convergence was verified similar and sa-
tisfactory for both considered flow-rates.
Fig. 6 shows values of averages (left) and RMSs (right) of the time
history of pressure coefficient for the three simulated diffuser geome-
tries, with 3 = 18 (a), 3 = 14 (b) and 3 = 10 (c). Statistics were
computed for the seven probes located at the mid-radius and mid-span
Fig. 5. Rotating and stationary channels of the pump, looking from the inlet.
of the diffuser vanes. They were ensemble-averaged over 10 impeller
The geometry is that of the original setting angle of the diffuser blades. The
revolutions. Here the pressure coefficient, cp, is defined as:
location of the pressure probes is identified by + symbols, in black for the
stationary diffuser probes and in white for the rotating impeller probes. The p p
elements of the pump shown in the figure are: clipped impeller blades (B); hub
cp = ,
0.5 u22 (5)
(C); diffuser blades (E); diffuser wall on the hub side (F); volute and casing
walls on the hub side (G). For limitation of space only the first and the last where p∞ is the pressure at the inlet section of the computational do-
impeller probes (moving from the pressure side towards the suction side) are main and ρ is the density of the fluid. In Fig. 6 the values of time-
explicitly indicated, as “r1” and “r7”. averaged pressure coefficient and RMSs of its fluctuations are reported,

48
A. Posa, A. Lippolis International Journal of Heat and Fluid Flow 75 (2019) 44–60

Fig. 6. Averages (left) and RMSs (right) of the time history of the pressure coefficient at the seven probes located at the mid-radius and mid-span of the diffuser vanes
for the off-design flow-rate: (a) 3 = 18 , (b) 3 = 14 and (c) 3 = 10 . In each chart the left, central and right columns stand for the results over the coarse, medium
and fine grids, respectively.

defined as follows:
Table 1
Error on the time-averaged pressure coefficients reported in Fig. 6 (left) on the N
(cp)i
i=1
coarse and medium grids, relative to the values computed on the fine grid. avg (cp) =
N
Averages across the seven probes at the diffuser mid span.
N
i=1
[(cp)i avg (cp)]2
Diffuser inlet angle Coarse grid Medium grid RMS (cp) =
N 1 (6)
3 = 18 1.24% 2.00%
= 14 0.48% 0.39% where N stands for the size of the time sample and (cp)i for an in-
3
3 = 10 1.45% 0.39% stantaneous value of the pressure coefficient. For the ensemble-aver-
aged statistics presented here the size of the sample for each case is
equal to the number of instantaneous realizations of the flow during the
computation of 10 impeller rotations, ranging from about 10,000 to
Table 2
about 20,000 from the coarse to the fine grids. The resolution in time
Error on the RMSs in time of pressure coefficient reported in Fig. 6 (right) on the
grows indeed as the resolution in space, in order to meet the stability
coarse and medium grids, relative to the values computed on the fine grid.
Averages across the seven probes at the diffuser mid span. requirements enforced by the CFL condition associated to the terms of
Eq. (1) handled by the explicit three-steps Runge-Kutta scheme. In
Diffuser inlet angle Coarse grid Medium grid
Tables 1 and 2 the average errors across the seven probes considered in
= 18 4.11% 4.06% Fig. 6 are reported for avg(cp) and RMS(cp), respectively, where the
3
= 14 7.70% 5.51% error was defined assuming as reference the values computed on the
3
3 = 10 5.75% 4.21% fine grid. For the first-order statistics in Fig. 6 (left) and Table 1 de-
pendence on grid resolution is very weak: differences across grids are

49
A. Posa, A. Lippolis International Journal of Heat and Fluid Flow 75 (2019) 44–60

Fig. 7. Phase-averaged fields of pressure coefficient at the off-design flow-rate. Location at the mid-span of the diffuser vanes. Angles 3 = 18 (a), 3 = 14 (b) and
3 = 10 (c). Results on the left, central and right columns from the coarse, medium and fine grids, respectively.

within 2% and even the coarse grid demonstrates to be suitable to the the stationary parts of the pump, providing details on the statistics of
computation of the time-averages of the pressure signals. For the the flow fields that are missing in the ensemble-averages. In both Figs. 7
second-order statistics of cp in Fig. 6 (right) and Table 2 results show a and 8 each row refers to a different diffuser setting angle (equivalent to
small dependence on grid resolution, which is more distinguishable 3 = 18 , 3 = 14 and 3 = 10 , from top to bottom). The fields on the
than for the time-averages, as expected, with errors within 6% on the left, central and right columns were generated by the computations on
medium grid and within 8% on the coarse grid. However, also for the the coarse, medium and fine grids, respectively. Both Figs. 7 and 8 refer
RMSs of pressure coefficient results can be considered satisfactory, to the location at the diffuser mid span. Results in Fig. 7 are practically
demonstrating that the computations carried out on the fine grid, uti- indistinguishable across resolutions, with increasing pressure rise from
lized for the analysis reported in the next sections, are able to capture (a) to (c), thanks to improving inflow conditions into the diffuser vanes,
the physics of the flow. as we will discuss later. Although the distribution of the phase-averaged
Comparisons across grids, dealing with the phase-averaged statistics RMSs of pressure coefficient is much more challenging to capture,
(synchronized with the impeller rotation), are reported in Figs. 7 and 8 especially at the interface between impeller and diffuser vanes, also in
for the averaged values of pressure coefficient and its RMSs, respec- Fig. 8 the agreement across grids is good. The topology of the areas
tively, again for the off-design flow-rate. Note that, although the defi- featuring the highest pressure fluctuations is very similar across the
nitions of average and RMS in time, reported in Eq. (6), still apply to the three adopted grids and the agreement is close also in quantitative
results shown in Figs. 7 and 8, phase-averages are a function of the terms. This is also demonstrated in both Table 3 and Table 4, where
position of the impeller blades. Therefore, whereas the sample utilized passage-averaged values of pressure coefficient and its RMSs are re-
for ensemble-averaging takes into account all instantaneous realiza- ported, respectively. Phase-averages, as those shown in both Figs. 7 and
tions of the flow during the simulation, after losing memory of the in- 8, were computed for twelve impeller positions, with an angular gap of
itial transient, for the phase-averages associated to a particular position 5° between consecutive configurations of the rotating channels. Note
of the impeller blades only the instantaneous solutions relative to that that the same impeller/diffuser configuration repeats every 60° of ro-
position are utilized for the computation of the statistics. This way tation, due to the periodicity of the impeller geometry. Then passage-
phase-averages allow us to isolate the coherence within the flow fields averages of cp and RMS(cp) were computed from phase-averages as the
and their dependence on the position of the moving parts, relative to arithmetic average across all twelve available phase-locked fields, as

50
A. Posa, A. Lippolis International Journal of Heat and Fluid Flow 75 (2019) 44–60

Fig. 8. Phase-averaged fields of pressure coefficient RMSs at the off-design flow-rate. Location at the mid-span of the diffuser vanes. Angles 3 = 18 (a), 3 = 14 (b)
and 3 = 10 (c). Results on the left, central and right columns from the coarse, medium and fine grids, respectively.

those shown in Figs. 7 and 8 for a particular phase-locked configura- 4.2. Effect of the working conditions
tion. In both Tables 3 and 4 results are reported as averages across the
diffuser inlet, the whole diffuser channels and the diffuser outlet. In In this section the influence on pressure fluctuations within rotating
addition, in Tables 3 and 4 the percentage error is provided in par- and stationary channels by the flow-rate conditions will be discussed,
entheses, assuming as reference the values computed on the fine grid. considering design and off-design flow-rates for the original diffuser
Table 3 points to the close agreement across grids in terms of passage- geometry. Fig. 9a reports the values of time-averaged RMSs for the
averaged cp, with errors always within 1%. Results show a good in- pressure coefficient at the seven probes located at the mid-span of the
dependence on grid resolution also in Table 4, although errors are diffuser vanes. Values of RMS(cp) are obviously higher at off-design
larger, due to the higher order of the statistics, compared to Table 3. In working conditions (right columns in Fig. 9a), since the flow is not
all cases featuring errors larger than 1% grid convergence is the ex- properly guided by the rotating channels into the diffuser vanes. The
pected one, while for lower errors it is likely that the size of the sta- interaction between moving and stationary parts is more unsteady, due
tistical sample plays a more significant role on the overall error. We to the incorrect incidence on the diffuser blades by the flow coming
would like to point that the present section was aimed mainly at de- from the impeller, leading to separation and back-flow phenomena
monstrating grid independence of the results that will be reported later (Posa et al., 2015; 2016a). As a consequence, RMSs of cp are roughly 2.5
in the manuscript, therefore more details on the flow physics will be times higher than those at the nominal flow-rate. Fig. 9a highlights also
discussed then. Nonetheless, it is interesting to see in Tables 3 and 4 that pressure fluctuations are not significantly dependent on the azi-
that at the off-design flow-rate a decrease of the angle β3 from 18° to 10° muthal location of the probes for both flow-rate conditions, at least at
leads to an increase of the values of pressure coefficient across the the considered radial location. From this result we can infer that there
diffuser channels from inlet to outlet and especially a sharp decrease of the effect of the asymmetry of the volute geometry is negligible. Fig. 9b
pressure fluctuations in time, demonstrating this way the beneficial provides another comparison between design and off-design working
effect of the modification of the setting angle of the blades over the conditions, where the spanwise evolution of RMS(cp) is shown within
performance of the centrifugal pump. All results presented in the next the diffuser vane facing the volute tongue (probes d1sβ, see Fig. 5 for
sections were taken from the computations performed on the fine grid. their position). At design flow-rate values are slightly higher on the

51
A. Posa, A. Lippolis International Journal of Heat and Fluid Flow 75 (2019) 44–60

Table 3 shroud side, where the flow through the rotating channels experiences
Passage-averaged values of pressure coefficient at the diffuser mid span, aver- higher pressure gradients and deviates more from the geometry of the
aged over the diffuser inlet, diffuser channels and diffuser outlet, respectively. impeller blades, leading to a slightly higher incidence on the diffuser
In parentheses the error relative to the values computed on the fine grid. Results blades. The impeller flow is better guided on the hub side and the
from the computations carried out at the off-design flow-rate.
diffuser inflow is therefore more properly aligned with the geometry of
Diffuser inlet the diffuser blades. Therefore, on the hub side the impeller/diffuser
interaction is smoother. However, at the design flow-rate gradients of
Diffuser inlet angle Coarse grid Medium grid Fine grid
RMS(cp) across the span keep almost negligible. They become more
3 = 18 0.7884 0.7844 0.7834
(0.65%) (0.13%) (0.00%)
obvious at off-design conditions, since the impeller outflow deviates
3 = 14 0.7944 0.7881 0.7889 more from the geometry of the diffuser inlet.
(0.70%) ( 0.11%) (0.00%) In Fig. 10 the spectra of the time history of cp at the probe d1s50 are
3 = 10 0.8236 0.8191 0.8216 shown at design (a) and off-design (b) working conditions, again for the
(0.25%) ( 0.30%) (0.00%) original orientation of the diffuser blades. Values on the horizontal axis
are normalized using the frequency of the impeller blades passage, fbp,
Diffuser channels
and are the same in the two panels of Fig. 10, for comparison purposes.
Diffuser inlet angle Coarse grid Medium grid Fine grid At the nominal flow-rate peaks occur at f = fbp and its higher harmo-
3 = 18 1.0477 1.0439 1.0516 nics, up to f = 7fbp , while energy at other frequencies is negligible. At
( 0.37%) ( 0.73%) (0.00%) the reduced flow-rate the main frequency is still fbp, but the role of the
3 = 14 1.0603 1.0561 1.0589 low frequencies is reinforced, with the maxima at the higher harmonics
(0.14%) ( 0.26%) (0.00%) of fbp much weaker and almost not distinguishable. This behavior of the
= 10 1.0861 1.0917 1.0930
3 spectra, verified across all diffuser vanes, is likely tied to the properties
( 0.63%) ( 0.12%) (0.00%)
of the flow coming from the impeller. We will show indeed in the next
Diffuser outlet section that even at smaller incidences, associated to smaller setting
angles of the diffuser blades, the higher harmonics of the blades passage
Diffuser inlet angle Coarse grid Medium grid Fine grid frequency keep almost negligible at the lower flow-rate. This is due to
1.1232 1.1193 1.1268
3 = 18 the weak influence of the orientation of the diffuser blades on the im-
( 0.32%) ( 0.66%) (0.00%)
peller outflow. At design flow-rate the impeller flow keeps attached,
= 14 1.1724 1.1694 1.1743
3
with a regular vortex shedding from the trailing edge of the rotating
( 0.17%) ( 0.42%) (0.00%)
= 10 1.2173 1.2283 1.2305 blades. The coherence of their wake is well distinguishable across the
3
( 1.07%) ( 0.18%) (0.00%) diffuser channels, up to the volute, affecting substantially the spectra of
the time history of pressure, dominated by the frequency of the blade
passage and its higher harmonics. Flow conditions within the diffuser
are significantly modified at the reduced flow-rate. A large region of
Table 4 separated flow is shed from the suction side of the impeller blades and
Passage-averaged values of RMS(cp) at the diffuser mid span, averaged over the the coherence of the impeller wake is lost very quickly within the dif-
diffuser inlet, diffuser channels and diffuser outlet, respectively. In parentheses fuser vanes, due to the lower level of organization of the flow coming
the error relative to the values computed on the fine grid. Results from the from the rotating channels, compared to that at design working con-
computations carried out at the off-design flow-rate. ditions, and the fast diffusion of its structures. At off-design working
Diffuser inlet conditions the energy associated to the fluctuations of cp is substantially
higher than at the nominal flow-rate, therefore two different vertical
Diffuser inlet angle Coarse grid Medium grid Fine grid scales were utilized in the two panels of Fig. 10. However, rotating stall
0.10751 0.11262 0.11283
3 = 18 was not observed from our computations. It was not reported also from
( 4.72%) ( 0.18%) (0.00%)
the measurements by Boccazzi et al. (2009). The highest peak of
= 14 0.07305 0.07527 0.07811
3
pressure fluctuations keeps at fbp, while typically rotating stall phe-
( 6.47%) ( 3.63%) (0.00%)
= 10 0.03054 0.03240 0.03434 nomena feature a very distinguishable lower frequency maximum,
3
( 11.05%) ( 5.63%) (0.00%) comparable or often higher than that at fbp (see, for instance, Sinha
et al., 2001; Zhang et al., 2015b and Pacot et al., 2016). We point also
Diffuser channels that, although a low frequency peak is visible in Fig. 10b, its detailed
analysis is beyond the scope of this work, especially due to computa-
Diffuser inlet angle Coarse grid Medium grid Fine grid
0.07063 0.07068 0.07038
tional cost considerations. Its study would require the advancement of
3 = 18
(0.35%) (0.43%) (0.00%) the solution over much longer times, because of the large period of the
3 = 14 0.04910 0.05070 0.05243 unsteady phenomena originating that peak. However, we verified in the
( 6.35%) ( 3.29%) (0.00%) spectra of the time-history of pressure coefficient at the impeller probes
3 = 10 0.02206 0.02241 0.02280 the presence of low frequency maxima even stronger than that in
( 3.24%) ( 1.71%) (0.00%) Fig. 10b, likely due to separation and back-flow phenomena at the
shroud (see Posa et al., 2016a for more details), featuring characteristic
Diffuser outlet
times that are longer than the period of the propeller rotation. Thus, we
Diffuser inlet angle Coarse grid Medium grid Fine grid can hypothesize that those pressure fluctuations are able to influence
3 = 18 0.03890 0.03780 0.03814 also the downstream diffuser channels. The same massive separation
(2.00%) ( 0.89%) (0.00%) phenomena do not occur at the nominal flow-rate, making the role of
3 = 14 0.02707 0.02760 0.02947 low frequencies substantially less significant within both impeller and
(0.00%)
( 8.13%) ( 6.33%) diffuser, as shown in Fig. 10a. Spectra similar to those in Fig. 10 were
= 10 0.01387 0.01387 0.01398
3
observed in all diffuser vanes at both nominal and reduced flow-rates,
( 0.75%) ( 0.77%) (0.00%)
therefore they were not reported here at additional tangential locations
within the diffuser.

52
A. Posa, A. Lippolis International Journal of Heat and Fluid Flow 75 (2019) 44–60

Fig. 9. RMSs of the time history of the pressure coefficient for the design and the off-design flow-rates (in each panel left and right columns, respectively): (a) at the
seven probes located at the mid-radius and mid-span of the diffuser vanes; (b) at the five probes located at the mid-radius of the diffuser vane facing the volute
tongue, with spanwise locations corresponding to 7%, 23%, 50%, 77% and 93% from the hub side. Results for the original diffuser geometry, with 3 = 18 .

In Fig. 11 the spectra of the time history of the pressure coefficient rate). The comparison between working conditions shows that on the
at the reduced flow-rate is reported at the diffuser probes located in the pressure side the sensitivity to flow-rate is rather limited, although cp is
same vane as in Fig. 10b, but at 23% and 77% of the span from the hub slightly lower in the off-design case. Instead, on the suction side de-
side. More energy at low frequencies, corresponding to stronger se- viations are larger, because of the wide region of recirculating flow
paration phenomena, was found on the shroud side (Fig. 11b), which is affecting the operation at reduced flow-rate. Actually, without separa-
consistent with the increase of the time-averaged RMS(cp) in the same tion pressure would grow monotonically from the suction side to the
area, shown in Fig. 9b. Nonetheless, the maximum at fbp does not ex- pressure side also in this case, keeping higher values across the whole
perience substantial variations across the span and the role of its higher azimuthal extent of the rotating channel, compared to the design con-
harmonics is confirmed negligible, in comparison with the spectrum dition, since a smaller flow-rate is processed by the impeller. In Fig. 13b
observed at the design working condition (Fig. 10a). The increased the RMSs of cp are also shown. They are higher at off-design across the
unsteady nature of the pressure fields within the diffuser vanes, moving whole azimuthal extent of the rotating channel, but differences between
from the hub to the shroud side, is shown in more details in Fig. 12, flow-rate conditions are much smaller than in the diffuser vanes, since
where phase-averaged RMSs of cp are reported for the off-design flow- they are originated mainly in the region of interface between rotating
rate at 23%, 50% and 77% of the diffuser span. It is evident that and stationary channels, where the highest pressure fluctuations occur,
pressure fluctuations within the stationary channels are lower on the as seen in Fig. 12. For both flow-rates RMSs in Fig. 13b are higher on
hub side. It is also obvious that the highest values populate the im- the pressure side, where the interaction with the downstream diffuser
peller/diffuser interface, being mainly associated to incorrect interac- blades is felt more directly, producing a periodic variation of the
tion between rotating and stationary blades (for additional details, see pressure field within the rotating channels.
Posa et al., 2015; Posa et al., 2016a).
The influence of the flow-rate conditions was studied also by means 4.3. Effect of the diffuser setting angle
of the pressure probes located within the impeller in the rotating re-
ference frame. Fig. 13a provides the time-averaged values of cp for As discussed above, the orientation of the diffuser blades was
Q = Qn and Q = Qr (left and right columns in Fig. 13a, respectively). At modified in order to decrease the incidence of the flow coming from the
the nominal flow-rate the evolution is monotonic from the pressure side rotating channels at the off-design condition. However, also the effect
towards the suction side, decreasing from the probe r1 to the probe r7, at the design flow-rate was analyzed. Posa and Lippolis (2018) pre-
as expected. In contrast, at the reduced flow-rate pressure reaches a sented a discussion mainly focused on the influence of the setting angle
plateau on the suction side, between probes r4 and r5, since a large on global performance, separation phenomena and velocity fields, in
steady separated area, moving with the impeller blade, populates the terms of both first and second order statistics. Here the analysis deals
suction side of the rotating channel (see Posa et al., 2015; Posa et al., with the impact on pressure fields, focusing especially on their unsteady
2016a for a discussion on separation phenomena at the reduced flow- features.

Fig. 10. Spectra of the time history of the pressure coefficient at the probe d1s50 for the design (a) and the off-design (b) flow-rates. In (a) and in (b) different scales
are adopted on the vertical axes. Results for the original diffuser geometry, with 3 = 18 .

53
A. Posa, A. Lippolis International Journal of Heat and Fluid Flow 75 (2019) 44–60

Fig. 11. Spectra of the time history of the pressure coefficient at the probe d1s23 (a) and the probe d1s77 (b) for the off-design flow-rate. Results for the original
diffuser geometry, with 3 = 18 .

Fig. 12. Phase-averaged fields of pressure coefficient RMSs at the off-design flow-rate. Results for the original diffuser geometry, with 3 = 18 . Spanwise locations at
23% (a), 50% (b) and 77% (c).

Fig. 14(top) reports the time-averaged values of cp at the probes where the smaller the angle β3, the larger the values of pressure within
located at the midspan of the diffuser vanes, dαs50, for the three si- the diffuser vanes, although the influence of the setting angle of the
mulated orientations of the blades. Fig. 14a refers to the design flow- diffuser blades is in this case less susbstantial than that observed at the
rate condition. The pressure rise through the pump is negatively af- nominal flow-rate.
fected by the modification of the diffuser geometry, causing high ne- Fig. 14(bottom) compares the RMSs of pressure coefficient for the
gative values of incidence at the leading edge of the blades, with con- three diffuser geometries. At design flow-rate a decrease of the angle β3
sequent separation phenomena on their pressure side. This effect is causes a substantial reduction of the pressure fluctuations at the diffuser
especially significant for 3 = 10 . The pressure drop from 3 = 14 to probes (see Fig. 14a), although the inflow into the diffuser is negatively
3 = 10 is indeed more obvious than that from 3 = 18 to 3 = 14 . affected by the increased (negative) incidence on the stationary blades.
Note also that, while the pressure rise is very uniform among the vanes This result can be actually explained taking into account that the
of the original diffuser geometry, its dependence on the azimuthal lo- modification of the diffuser geometry leads to an increased radial gap
cation of the probe becomes more evident at reduced setting angles. As between rotating and stationary parts, with a smoother time evolution
expected, at the lower flow-rate the effect of the modification of the of the pressure field. Furthermore, we have verified that, although an
diffuser geometry is instead beneficial, as shown in Fig. 14b (top), incorrect incidence at reduced setting angles originates an increase of

Fig. 13. Averages (a) and RMSs (b) of the time history of the pressure coefficient for the design and the off-design flow-rates (in each panel left and right columns,
respectively) at the seven probes located within one of the impeller vanes. Evolution from the pressure side (r1) towards the suction side (r7). Results for the original
diffuser geometry, with 3 = 18 .

54
A. Posa, A. Lippolis International Journal of Heat and Fluid Flow 75 (2019) 44–60

Fig. 14. Averages (top) and RMSs (bottom) of the time history of the pressure coefficient for the design (a) and the off-design (b) flow-rates at the seven probes
located at the mid-radius and mid-span of the diffuser vanes: 3 = 18 (left columns); 3 = 14 (central columns); 3 = 10 (right columns).

pressure fluctuations in some areas of the diffuser vanes (their pressure at the location of the diffuser probes, that are indeed higher than those
side), the present diffuser probes are located closer to the suction side, at the design flow-rate. Lower setting angles improve the diffuser in-
where the effect of the incorrect inflow is less dramatic. Also in Fig. 14b flow, decreasing substantially the gap between pressure fluctuations at
(bottom), dealing with the off-design flow-rate, a smoother interaction design and reduced flow-rates, as shown by the comparison between
impeller/diffuser at smaller setting angles leads to lower values of RMS the two bottom panels of Fig. 14.
(cp), thanks to both increased impeller/diffuser gap and reduced in- The beneficial effect of a smaller setting angle of the diffuser blades
cidence. For the original diffuser geometry the large positive incidence at reduced flow-rate is shown in more detail via the fields in Fig. 15,
triggers separation on the suction side, enhancing pressure fluctuations where the phase-averaged pressure coefficient and its RMSs are

Fig. 15. Fields of phase-averages (top) and phase-averaged RMSs (bottom) of pressure coefficient at the off-design flow-rate. Location at the mid-span of the diffuser
vanes. Setting angles 3 = 18 (a), 3 = 14 (b) and 3 = 10 (c).

55
A. Posa, A. Lippolis International Journal of Heat and Fluid Flow 75 (2019) 44–60

Fig. 16. Spectra of the time history of the pressure coefficient at the probe d1s50 for the design (top) and off-design (bottom) flow-rates with the modified diffuser
geometry: (a) 3 = 14 ; (b) 3 = 10 .

presented at the mid-span location for the three simulated diffuser noting that the choice of the vertical scales in Fig. 16 was indeed based
geometries. In Fig. 15 (top) the pressure rise through the impeller on comparison purposes. At the nominal flow-rate in Fig. 16 (top) the
channels is practically unaffected by the setting angle of the diffuser maxima corresponding to the frequency of the blade passage and its
blades. This result will be discussed in more detail later, looking at the higher harmonics experience a dramatic decrease, consistent with the
pressure probes within the impeller. Differences across values of β3 are discussion above. However, they are still very well distinguishable, at
instead distinguishable within the diffuser, where the pressure rise least up to f = 4fbp , with a limited role of the low frequencies, in-
becomes gradually larger from 3 = 18 to 3 = 10 , showing that the dicating that pressure fluctuations keep almost synchronized with the
trend seen in Fig. 14b (top) is not limited to the particular location of rotation of the impeller blades and are not significantly affected by
the probes placed within the diffuser and involves the whole stationary important slow unsteady phenomena. At the lower flow-rate in Fig. 16
vanes and the downstream volute channel. Cross-stream gradients (bottom) the peak at the frequency fbp fades out according to a factor
within both impeller and diffuser display growing values of pressure about equal to 3 from 3 = 18 to 3 = 10 , being originally affected by
from the concave to the convex sides of both rotating and stationary the unsteady separation of the flow over the suction side of the leading
vanes. In agreement with Fig. 14b (bottom), referring to ensemble- edge of the stationary blades, induced by the periodic passage of the
averages, smaller setting angles allow a decrease of the phase-averaged impeller blades. Also the energy at the low frequencies experiences a
pressure fluctuations in Fig. 15 (bottom). Lower values of RMS(cp) are substantial decrease, thanks to a sharp reduction of the unsteady large-
already distinguishable in Fig. 15b (bottom, 3 = 14 ), compared to scale back-flow phenomena produced during the operation of the ori-
Fig. 15a (bottom, 3 = 18 ), both at the impeller/diffuser interface and ginal diffuser at the considered low flow-rate, as discussed by Posa et al.
through the downstream diffuser channels, especially on their suction (2015, 2016a).
side. This result is tied to both reduced values of incidence and an in- Some details relative to the evolution of the pressure field over the
creased gap between impeller trailing edges and diffuser leading edges. span of the diffuser vanes are also provided. Results for the time-
In contrast, pressure fluctuations on the suction side of the rotating averaged values of cp (for brevity not reported here) show that actually
channels do not experience substantial variations, being mainly tied to the pressure rise through the diffuser vanes is rather uniform across the
separation phenomena almost unaffected by impeller/diffuser interac- span and the effect of the variation of the diffuser geometry is similar at
tion (see Posa et al., 2015 and Posa and Lippolis, 2018 for details). all monitored spanwise locations. The behavior of the pressure fluc-
Instead, in the downstream region of the pressure side of the rotating tuations is more interesting. As shown in Fig. 17a, at the nominal flow-
channels fluctuations undergo a decrease from Fig. 15a (bottom) to rate they keep slightly higher on the shroud side of the diffuser vanes.
Fig. 15b (bottom), being mainly tied to the unsteady phenomena at the In all cases near the shroud a thicker boundary layer develops on the
impeller/diffuser interface. However, such decrease is far more dra- suction side of the impeller blades, affecting the diffuser channels with
matic in Fig. 15c (bottom), referring to 3 = 10 . Again, pressure fluc- a wider impeller wake. At all spanwise locations a significant reduction
tuations do not change substantially within the impeller channels, of RMS(cp) occurs from the setting angle 3 = 18 to the setting angle
whereas the effect of the smoother interaction between moving and 3 = 14 , while changes between 3 = 14 and 3 = 10 are rather lim-
stationary parts is very evident from the impeller outflow up to the ited. At the reduced flow-rate (Fig. 17b) the spanwise evolution be-
discharge channel downstream of the volute. comes more uniform from 3 = 18 to 3 = 10 , coupled with a dramatic
Spectra of cp time history at the probe d1s50 are provided in Fig. 16, reduction of pressure fluctuations. In Fig. 17b we can see that the re-
for the design (top) and the off-design (bottom) flow-rates, respectively. duction of pressure fluctuations for decreasing values of the angle β3 is
They refer to the modified orientations of the diffuser blades ( 3 = 14 monotonic, in contrast with the plateau achieved between 3 = 14 and
and 3 = 10 ) and need to be compared with those in Fig. 10. It is worth 3 = 10 at the design flow-rate in Fig. 17a. Maybe this behavior is due

56
A. Posa, A. Lippolis International Journal of Heat and Fluid Flow 75 (2019) 44–60

Fig. 17. RMSs in time of pressure coefficient for the design (a) and the off-design (b) flow-rates at the five probes located at the diffuser mid-radius along the span of
the vane facing the volute tongue: 3 = 18 (left columns); 3 = 14 (central columns); 3 = 10 (right columns).

Fig. 18. Phase-averaged fields of pressure coefficient RMSs at the off-design flow-rate. Spanwise locations at 23% (a), 50% (b) and 77% (c). Setting angles corre-
sponding to 3 = 14 (top) and 3 = 10 (bottom).

to the location of the diffuser probes, near the suction side of the blades. the impeller/diffuser interface, but overall differences across the span
At off-design a gradual improvement of incidence conditions allows for are substantially smaller, compared to both cases with 3 = 18 and
a gradual decrease of pressure fluctuations, thanks to a smoother dif- 3 = 14 .
fuser inflow. At design flow-rate, although more unfavorable diffuser The influence of the setting angle of the diffuser blades was verified
inflow conditions are produced for 3 = 14 , in comparison with significantly more marginal on pressure signals at the impeller probes
3 = 18 , lower pressure fluctuations are also observed, tied to an in- rγ. Fig. 19 (top) shows the time-averaged values of cp for the three si-
creased radial gap between impeller and diffuser blades, whereas the mulated diffuser geometries. In Fig. 19a (top) the case at the nominal
influence of a further increase of such gap from 3 = 14 to 3 = 10 flow-rate is considered. Obviously the effect on the pressure rise
seems less significant. Actually, albeit at the nominal flow-rate the in- through the impeller is negligible. A similar result is visible at the re-
cidence angle over the leading edge of the diffuser blades grows for duced flow-rate in Fig. 19b (top). This is consistent with the discussion
decreasing values of the angle β3, it is negative (Posa and reported above, where we pointed to the limited influence of the or-
Lippolis, 2018), generating an increase of the pressure fluctuations ientation of the diffuser blades on the flow through the rotating chan-
mainly in the area of the pressure side of the diffuser vanes, far away nels, with the exception of their outlet.
from the location of the probes placed within the stationary channels. The RMSs of the time history of cp at the same impeller probes are
More details on pressure fluctuations are provided in Fig. 18, to be shown in Fig. 19 (bottom), where the dependence on the orientation of
compared to the spanwise evolution of the phase-averaged fields of the diffuser blades is more distinguishable. For the nominal flow-rate,
RMS(cp) reported in Fig. 12 for the original diffuser angle 3 = 18 . For considered in Fig. 19a (bottom), a reduction of the angle β3 does not
the case 3 = 14 in Fig. 18 (top) pressure fluctuations are still sig- affect the suction side, but lower pressure fluctuations occur on the
nificantly higher on the shroud side, with the highest values populating pressure side. The distance between trailing edge of impeller blades and
the area between impeller and diffuser vanes and affecting especially leading edge of diffuser blades increases, decreasing the pulsations of
the upstream region of the latter. This is not the case for 3 = 10 in pressure associated to their interaction. As a consequence of decreased
Fig. 18 (bottom). Values on the shroud side of the diffuser channels are impeller/diffuser interaction, also gradients of RMS(cp) between pres-
still slightly higher at the leading edge of the stationary blades and at sure and suction sides become smaller, with closer values across the

57
A. Posa, A. Lippolis International Journal of Heat and Fluid Flow 75 (2019) 44–60

Fig. 19. Averages (top) and RMSs (bottom) of the time history of the pressure coefficient for the design (a) and the off-design (b) flow-rates at the seven probes
located within one of the impeller vanes from the pressure side (r1) towards the suction side (r7): 3 = 18 (left columns); 3 = 14 (central columns); 3 = 10 (right
columns).

seven impeller probes. At the reduced flow-rate, whose results are component in the spectra, even at the reduced flow-rate. In contrast, the
shown in Fig. 19b (bottom), the impact of the modified diffuser geo- role of its higher harmonics was found negligible, compared to the si-
metry is more evident across the whole azimuth, since at this working mulations performed at the design flow-rate. The likely reason is that at
condition the impeller/diffuser interaction is much stronger for the nominal conditions coherence within the diffuser vanes is dominated by
original diffuser design. Therefore, the beneficial effect of lower values the wake of the impeller blades, keeping well distinguishable up to the
of β3 is felt even on the suction side of the impeller blades, although it outlet of the stationary channels. This is not the case at the lower flow-
does not face directly the stationary channels. However, the depen- rate, due to incorrect inflow, as discussed in detail by Posa et al. (2015,
dence of pressure fluctuations at the impeller probes on the orientation 2016a). More significant gradients across the span of the diffuser were
of the diffuser blades keeps substantially weaker than that seen above found at off-design, with higher fluctuations on the shroud side, com-
within the diffuser. pared to the hub side, due to larger streamwise pressure gradients ex-
perienced by the flow coming from the impeller channels.
5. Conclusions Smaller inlet angles of the diffuser blades allowed a higher pressure
rise within the diffuser channels at the lower flow-rate, while the op-
In this paper Large Eddy Simulation, coupled with an Immersed posite occurred at the nominal flow-rate. This effect was associated to
Boundary methodology, was utilized to study pressure fluctuations reduced incidence in the former case and increased incidence in the
within a small radial pump. LES is especially suitable to this problem, latter case. However, pressure fluctuations at the diffuser probes were
due to its eddy-resolving nature. The present approach was validated in substantially lower at both flow-rates, due to the larger radial gap be-
our earlier studies at both design and off-design working conditions tween impeller and diffuser blades, with a weaker interaction between
(Posa et al., 2011; 2015; 2016a) for the original diffuser geometry, by moving and stationary elements. Such weaker interaction was also
comparison with the PIV experiments by Boccazzi et al. (2009). A evident in the spectra of the pressure time history, with lower maxima
general discussion about the dependence of the flow physics on the at the blades passing frequency and its harmonics. At the reduced flow-
setting angle of the diffuser blades was presented via velocity statistics rate smaller spanwise gradients of pressure fluctuations were also
by Posa and Lippolis (2018). Pressure signals were recorded at probes found, compared to the original diffuser geometry. The impact of the
located within both impeller and diffuser vanes. Here their sensitivity to variation of the setting angle of the diffuser blades was more marginal
flow-rate and diffuser geometry was investigated via both ensemble- on the impeller, where pressure rise was almost unaffected. Pressure
averaged and phase-averaged statistics. A grid refinement study was fluctuations were reduced mainly on the pressure side of the impeller
also reported in this study, demonstrating a satisfactory independence blades, facing directly the downstream diffuser channels, although at
of both first and second order statistics from grid resolution. Results the off-design flow-rate a decrease of pressure RMSs on the suction side
have shown that pressure fluctuations within both impeller and diffuser was also distinguishable, thanks to a smoother transition of the flow
were higher at reduced flow-rate, compared to the design condition. from the impeller to the diffuser. It is worth noting that the present
However, this effect was much more obvious at the impeller/diffuser analysis involves both a change of the incidence angle at the leading
interface and within the stationary channels, especially in their up- edge of the diffuser blades and an increased radial gap between rotating
stream region, where pressure RMSs were about 2.5 times higher at off- and stationary parts. This choice, with a rotation of the diffuser blades
design, while within the impeller the increase was substantially milder. around their mid camber location, was considered more practical for an
Nonetheless, rotating stall was not detected within the diffuser channels actual machine. However, in the future it would be interesting to ex-
and the frequency of the passage of the impeller blades kept the main plore a modification of the diffuser geometry able to isolate each effect.

58
A. Posa, A. Lippolis International Journal of Heat and Fluid Flow 75 (2019) 44–60

Acknowledgments ASME J. Fluids Eng. 138 (11), 111102. https://doi.org/10.1115/1.4033423.


Parrondo-Gayo, J., González-Pérez, J., Fernández-Francos, J., 2002. The effect of the
operating point on the pressure fluctuations at the blade passage frequency in the
The authors are grateful to A. Boccazzi and R. Miorini for providing volute of a centrifugal pump. ASME J. Fluids Eng. 124 (3), 784–790. https://doi.org/
their experimental results. 10.1115/1.1493814.
Pei, J., Yuan, S., Benra, F.-K., Dohmen, H., 2012. Numerical prediction of unsteady
pressure field within the whole flow passage of a radial single-Blade pump. ASME J.
References Fluids Eng. 134 (10), 101103. https://doi.org/10.1115/1.4007382.
Peskin, C., 1977. Numerical analysis of blood flow in the heart. J. Comput. Phys. 25 (3),
Arndt, N., Acosta, A., Brennen, C., Caughey, T., 1990. Experimental investigation of rotor- 220–252. https://doi.org/10.1016/0021-9991(77)90100-0.
Stator interaction in a centrifugal pump with several vaned diffusers. ASME J. Peskin, C., 2002. The immersed boundary method. Acta Numerica 11, 479–517. https://
Turbomach. 112 (1), 98–108. https://doi.org/10.1115/1.2927428. doi.org/10.1017/S0962492902000077.
Balaras, E., 2004. Modeling complex boundaries using an external force field on fixed Posa, A., Balaras, E., 2016. A numerical investigation of the wake of an axisymmetric
cartesian grids in large-Eddy simulations. Comput. Fluids 33 (3), 375–404. https:// body with appendages. J. Fluid Mech. 792, 470–498. https://doi.org/10.1017/jfm.
doi.org/10.1016/S0045-7930(03)00058-6. 2016.47.
Balaras, E., Schroeder, S., Posa, A., 2015. Large-Eddy simulations of submarine propellers. Posa, A., Balaras, E., 2018. Large eddy simulation of an isolated vertical axis wind tur-
J. Ship Res. 59 (4), 227–237. https://doi.org/10.5957/JOSR.59.4.150047. bine. J. Wind Eng. Ind. Aerodyn. 172, 139–151. https://doi.org/10.1016/j.jweia.
Barrio, R., Parrondo, J., Blanco, E., 2010. Numerical analysis of the unsteady flow in the 2017.11.004.
near-tongue region in a volute-type centrifugal pump for different operating points. Posa, A., Balaras, E., 2018. Large-Eddy simulations of a notional submarine in towed and
Comput. Fluids 39 (5), 859–870. https://doi.org/10.1016/j.compfluid.2010.01.001. self-propelled configurations. Comput. Fluids 165, 116–126. https://doi.org/10.
Barsky, D., Posa, A., Rahromostaqim, M., Leftwich, M.C., Balaras, E., 2014. Experimental 1016/j.compfluid.2018.01.013.
and computational wake characterization of a vertical axis wind turbine. AVIATION Posa, A., Lippolis, A., 2018. A LES investigation of off-design performance of a centrifugal
2014, 32nd AIAA Applied Aerodynamics Conference, 16–20 June 2014, Atlanta, pump with variable-geometry diffuser. Int. J. Heat Fluid Flow 70, 299–314. https://
USA. https://doi.org/10.2514/6.2014-3141. doi.org/10.1016/j.ijheatfluidflow.2018.02.011.
Boccazzi, A., Miorini, R., Sala, R., Marinoni, F., 2009. Unsteady flow field in a radial Posa, A., Lippolis, A., Balaras, E., 2015. Large-Eddy simulation of a mixed-flow pump at
pump vaned diffuser. 8th European Conference on Turbomachinery: Fluid Dynamics off-design conditions. ASME J. Fluids Eng. 137 (10), 101302. https://doi.org/10.
and Thermodynamics, ETC 2009 - Conference Proceedings. pp. 1103–1112. 1115/1.4030489.
Chorin, A., 1967. The numerical solution of the Navier–Stokes equations for an in- Posa, A., Lippolis, A., Balaras, E., 2016. Investigation of separation phenomena in a radial
compressible fluid. Bull. Am. Math. Soc. 73 (6), 928–931. pump at reduced flow rate by large eddy simulation. ASME J. Fluids Eng. 138 (12),
Ducros, F., Comte, P., Lesieur, M., 1996. Large-eddy simulation of transition to turbulence 121101. https://doi.org/10.1115/1.4033843.
in a boundary layer developing spatially over a flat plate. J. Fluid Mech. 326, 1–36. Posa, A., Lippolis, A., Verzicco, R., Balaras, E., 2011. Large-Eddy simulations in mixed-
https://doi.org/10.1017/S0022112096008221. flow pumps using an immersed-boundary method. Comput. Fluids 47 (1), 33–43.
Ducros, F., Nicoud, F., Poinsot, T., 1998. Wall-adapting local eddy-viscosity models for https://doi.org/10.1016/j.compfluid.2011.02.004.
simulations in complex geometries. Numer. Methods Fluid Dyn. VI 293–299. Posa, A., Parker, C.M., Leftwich, M.C., Balaras, E., 2016. Wake structure of a single
Feng, J., Benra, F.-K., Dohmen, H., 2007. Numerical investigation on pressure fluctuations vertical axis wind turbine. Int. J. Heat Fluid Flow 61, 75–84. https://doi.org/10.
for different configurations of vaned diffuser pumps. Int. J. Rotating Mach. 2007, 1016/j.ijheatfluidflow.2016.02.002.
34752. https://doi.org/10.1155/2007/34752. Rossi, T., Toivanen, J., 1999. Parallel fast direct solver for block tridiagonal systems with
Fu, D., Wang, F., Zhou, P., Xiao, R., Yao, Z., 2018. Impact of impeller stagger angles on separable matrices of arbitrary dimension. SIAM J. Scient. Comput. 20 (5),
pressure fluctuation for a double-suction centrifugal pump. Chin. J. Mech. Eng. 31 1778–1793. https://doi.org/10.1137/S1064827597317016.
(1), 10. https://doi.org/10.1186/s10033-018-0203-z. Shi, F., Tsukamoto, H., 2001. Numerical study of pressure fluctuations caused by im-
Fu, Y., Yuan, J., Yuan, S., Pace, G., D’Agostino, L., Huang, P., Li, X., 2015. Numerical and peller-diffuser interaction in a diffuser pump stage. ASME J. Fluids Eng. 123 (3),
experimental analysis of flow phenomena in a centrifugal pump operating under low 466–474. https://doi.org/10.1115/1.1385835.
flow rates. ASME J. Fluids Eng. 137 (1), 011102. https://doi.org/10.1115/1. Si, Q., Yuan, J., Yuan, S., Wang, W., Zhu, L., Bois, G., 2014. Numerical investigation of
4027142. pressure fluctuation in centrifugal pump volute based on SAS model and experi-
Gao, B., Zhang, N., Li, Z., Ni, D., Yang, M., 2016. Influence of the blade trailing edge mental validation. Adv. Mech. Eng. 2014 (6), 972081. https://doi.org/10.1155/
profile on the performance and unsteady pressure pulsation in a low specific speed 2014/972081.
centrifugal pump. ASME J. Fluids Eng. 138 (5), 051106. https://doi.org/10.1115/1. Si, Q., Yuan, S., Yuan, J., Liang, Y., 2013. Investigation on flow-Induced noise due to
4031911. backflow in low specific speed centrifugal pumps. Adv. Mech. Eng. 2013, 109048.
Gao, Z., Zhu, W., Lu, L., Deng, J., Zhang, J., Wuang, F., 2014. Numerical and experimental https://doi.org/10.1155/2013/109048.
study of unsteady flow in a large centrifugal pump with stay vanes. ASME J. Fluids Sinha, M., Pinarbasi, A., Katz, J., 2001. The flow structure during onset and developed
Eng. 136 (7), 071101. https://doi.org/10.1115/1.4026477. states of rotating stall within a vaned diffuser of a centrifugal pump. ASME J. Fluids
Kato, C., Kaiho, M., Manabe, A., 2003. An overset finite-element large-Eddy simulation Eng. 123 (3), 490–499. https://doi.org/10.1115/1.1374213.
method with applications to turbomachinery and aeroacoustics. ASME J. Appl. Mech. Smagorinsky, J., 1963. General circulation experiments with the primitive equations: i.
70 (1), 32–43. https://doi.org/10.1115/1.1530637. The basic experiment. Mon. Weather Rev. 91 (3), 99–164. https://doi.org/10.1175/
Kato, C., Yamade, Y., Wang, H., Guo, Y., Miyazawa, M., Takaishi, T., Yoshimura, S., 1520-0493(1963)091<0099:GCEWTP>2.3.CO;2.
Takano, Y., 2007. Numerical prediction of sound generated from flows with a low Spence, R., Amaral-Teixeira, J., 2008. Investigation into pressure pulsations in a cen-
Mach number. Comput. Fluids 36 (1), 53–68. https://doi.org/10.1016/j.compfluid. trifugal pump using numerical methods supported by industrial tests. Comput. Fluids
2005.07.006. 37 (6), 690–704. https://doi.org/10.1016/j.compfluid.2007.10.001.
Kato, C., Yoshimura, S., Yamade, Y., Jiang, Y., Wang, H., Imai, R., Katsura, H., Yoshida, Spence, R., Amaral-Teixeira, J., 2009. A CFD parametric study of geometrical variations
T., Takano, Y., 2005. Prediction of the noise from a multi-stage centrifugal pump. on the pressure pulsations and performance characteristics of a centrifugal pump.
ASME 2005 Fluids Engineering Division Summer Meeting, Houston, Texas, USA, June Comput. Fluids 38 (6), 1243–1257. https://doi.org/10.1016/j.compfluid.2008.11.
1923, 2005. 1. pp. 1273–1280. https://doi.org/10.1115/FEDSM2005-77312. 013.
Kye, B., Park, K., Choi, H., Lee, M., Kim, J., 2018. Flow characteristics in a volute-type Swarztrauber, P., 1974. A direct method for the discrete solution of separable elliptic
centrifugal pump using large eddy simulation. Int. J. Heat Fluid Flow 72, 52–60. equations. SIAM J. Numer. Anal. 11 (6), 1136–1150. https://doi.org/10.1137/
https://doi.org/10.1016/j.ijheatfluidflow.2018.04.016. 0711086.
Kye, B., Park, K., Choi, H., Lee, M., Kim, J.H., 2017. Flow characteristics in a centrifugal Van Kan, J., 1986. A second-order accurate pressure-correction scheme for viscous in-
pump using large eddy simulation. 10th International Symposium on Turbulence and compressible flow. SIAM J. Scient. Stat. Comput. 7 (3), 870–891. https://doi.org/10.
Shear Flow Phenomena, TSFP 2017, Chicago, IL, USA, 6–9 July 2017. 1137/0907059.
Lee, J., Choi, H., Park, N., 2010. Dynamic global model for large Eddy simulation of Vanella, M., Posa, A., Balaras, E., 2014. Adaptive mesh refinement for immersed
transient flow. Phys. Fluids 22 (7), 075106. https://doi.org/10.1063/1.3459156. boundary methods. ASME J. Fluids Eng. 136 (4), 040909. https://doi.org/10.1115/1.
Liu, M., Tan, L., Cao, S., 2018. Influence of geometry of inlet guide vanes on pressure 4026415.
fluctuations of a centrifugal pump. ASME J. Fluids Eng. 140 (9), 091204. https://doi. Verzicco, R., Mohd-Yusof, J., Orlandi, P., Haworth, D., 2000. Large eddy simulation in
org/10.1115/1.4039714. complex geometric configurations using boundary body forces. AIAA J, 38 (3),
Mittal, R., Iaccarino, G., 2005. Immersed boundary methods. Annu. Rev. Fluid Mech. 37, 427–433. https://doi.org/10.2514/2.1001.
239–261. https://doi.org/10.1146/annurev.fluid.37.061903.175743. Wang, W., Wang, Y., 2013. Analysis of inner flow in low specific speed centrifugal pump
Nicoud, F., Ducros, F., 1999. Subgrid-scale stress modelling based on the square of the based on LES. J. Mech. Sci. Technol. 27 (6), 1619–1626. https://doi.org/10.1007/
velocity gradient tensor. Flow, Turbul. Combust. 62 (3), 183–200. https://doi.org/ s12206-013-0408-0.
10.1023/A:1009995426001. Yang, J., Balaras, E., 2006. An embedded-Boundary formulation for large-Eddy simula-
Orlanski, I., 1976. A simple boundary condition for unbounded hyperbolic flows. J. tion of turbulent flows interacting with moving boundaries. J. Comput. Phys. 215 (1),
Comput. Phys. 21 (3), 251–269. https://doi.org/10.1016/0021-9991(76)90023-1. 12–40. https://doi.org/10.1016/j.jcp.2005.10.035.
Pacot, O., Kato, C., Guo, Y., Yamade, Y., Avellan, F., 2016. Large eddy simulation of the Yang, J., Pavesi, G., Yuan, S., Cavazzini, G., Ardizzon, G., 2015. Experimental char-
rotating stall in a pump-turbine operated in pumping mode at a part-load condition. acterization of a pump-turbine in pump mode at hump instability region. ASME J.

59
A. Posa, A. Lippolis International Journal of Heat and Fluid Flow 75 (2019) 44–60

Fluids Eng. 137 (5), 051109. https://doi.org/10.1115/1.4029572. Zhang, N., Yang, M., Gao, B., Li, Z., Ni, D., 2015. Experimental investigation on unsteady
Yao, Z., Wang, F., Qu, L., Xiao, R., He, C., Wang, M., 2011. Experimental investigation of pressure pulsation in a centrifugal pump with special slope volute. ASME J. Fluids
time-frequency characteristics of pressure fluctuations in a double-suction centrifugal Eng. 137 (6), 061103. https://doi.org/10.1115/1.4029574.
pump. ASME J. Fluids Eng. 133 (10), 101303. https://doi.org/10.1115/1.4004959. Zheng, L., Dou, H., Chen, X., Zhu, Z., Cui, B., 2018. Pressure fluctuation generated by the
Zhang, N., Yang, M., Gao, B., Li, Z., Ni, D., 2015. Experimental and numerical analysis of interaction of blade and tongue. J. Therm. Sci. 27 (1), 8–16. https://doi.org/10.
unsteady pressure pulsation in a centrifugal pump with slope volute. J. Mech. Sci. 1007/s11630-018-0978-3.
Technol. 29 (10), 4231–4238. https://doi.org/10.1007/s12206-015-0919-y.

60

You might also like