You are on page 1of 10

Proceedings of ASME Turbo Expo 2016: Turbomachinery Technical Conference and Exposition

GT2016
June 13 – 17, 2016, Seoul, South Korea

GT2016-56036

NUMERICAL STUDY OF THE REYNOLDS NUMBER EFFECT ON THE


CENTRIFUGAL COMPRESSOR PERFORMANCE AND LOSSES

Jonna Tiainen∗, Ahti Jaatinen-Värri, Aki Grönman, Jari Backman


Laboratory of Fluid Dynamics
LUT School of Energy Systems
Lappeenranta University of Technology
53851 Lappeenranta, Finland
Email: jonna.tiainen@lut.fi

ABSTRACT NOMENCLATURE
The efficiency is reduced in very small centrifugal compres- Latin alphabet
sors due to low Reynolds numbers. In the past, the effect of a fraction of Reynolds-independent losses [-]
the Reynolds number on centrifugal compressor performance b fraction of Reynolds-dependent losses [-]
has been studied experimentally, and empirical correction equa- b2 blade height [m]
tions for the efficiency have been derived based on those results. Bref coefficient in Eqns. (4) and (5) [-]
There is a lack of numerical investigations into the effect of the c absolute velocity [m/s]
Reynolds number on centrifugal compressor performance and c chord length [m]
losses. This paper aims to compare the numerical results to the c coefficient in Eqn. (2) [-]
efficiencies predicted by the correction equations found in the cf friction coefficient [-]
literature. The loss generation in the impeller blade passages is D diameter [m]
also studied in order to find out which loss production mecha- f friction factor [-]
nism has the most potential to be reduced or eliminated. h specific enthalpy [J/kg]
U2
The effect of the Reynolds number on compressor perfor- MaU Mach number based on tip speed, MaU = √γRT [-]
1
mance is investigated in the chord Reynolds number range vary- N rotational speed [rpm]
ing from 0.8 · 105 to 17 · 105 by simulating numerically the orig- n Reynolds-number-ratio exponent in Eqns. (1) and (3) [-]
inal compressors and downscaled ones. The numerical results p pressure [Pa]
are validated against experimental data and the results are com- qm mass flow rate [kg/s]
pared with the efficiency correction equations used in the liter- qv volume flow rate [m3 /s]
ature. The results indicate that the performance of the down- R specific gas constant [J/kgK]
scaled compressors follow quite precisely the most recently pub- Reb2 Reynolds number based on blade height, Reb2 = Uν21b2 [-]
lished correction equation. The results also show that the in- Rec Reynolds number based on chord length, Rec = wν11c [-]
creased losses in low-Reynolds-number compressors are caused
ReD2 Reynolds number based on tip diameter, ReD2 = U2νD1 2 [-]
both by the relatively increased boundary layer thickness and by
T temperature [K]
the shear stress resulting from the increased vorticity.
t tip clearance [m]
U tip speed [m/s]
w relative velocity [m/s]
∗ Address all correspondence to this author.

1 Copyright © 2016 by ASME

Downloaded From: https://proceedings.asmedigitalcollection.asme.org on 06/30/2019 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Greek alphabet and Reynolds-independent losses are not accounted for at all.
η isentropic total-to-total efficiency [-] In the correction equations, different definitions for the
γ ratio of specific heats [-] Reynolds number (Reb2 , Rec , and ReD2 ) and for the friction factor
µ0 work input coefficient [-] (cf and f ) are used. Some of the equations are not based on the
ν kinematic viscosity [m2 /s] change in the friction factor but on the change in the Reynolds
φ flow coefficient, φ = UqDv 2 [-] number. According to Casey and Robinson [1], the correction
2 2
π pressure ratio [-] equations, which are based on the change in the friction fac-
ψ pressure coefficient, ψ = ∆hs
[-] tor, are recommended by most authors because the friction fac-
U22
tor accounts for the effect of both the surface roughness and the
Abbreviations
Reynolds number.
FB full blade
The improvement in Casey’s correction equations (2), (4),
PS pressure side
and (5) compared to the old empirical formula (1) is that Casey’s
SB splitter blade
D2,scaled equations do not take the Reynolds-independent losses into ac-
SF diameter scaling factor, SF = D2,baseline count as a constant fraction but as a constant amount, because the
SS suction side fraction is not constant with a changing Reynolds number. The
Subscripts difference between Casey’s original equation (2) and the modi-
1 impeller inlet fied one (5) is that the original equation is based on the friction
2 impeller outlet factor of fully developed pipe flow and the modified one is based
m meridional on the flat plate friction factor. Casey and Robinson [1] state that
ref reference ”the blockage due to the boundary layers may become as high
s isentropic, static as 10%, but fully developed flow with merged boundary layers
tt total-to-total from both sides of the flow channels does not usually occur at the
u tangential design point”.
The slope of Pelz and Stonjek’s correction equation (6) is
more negative than that of Dietmann and Casey because Pelz and
INTRODUCTION Stonjek ignore the Reynolds-independent losses. By ignoring
In the past, the effect of the Reynolds number on centrifugal Reynolds-independent losses, Pelz and Stonjek [5] assume that
compressor performance has been studied experimentally and there are only losses due to friction. According to Dietmann and
empirical correction equations for the efficiency were derived Casey [3], in addition to frictional losses there are also shock,
based on those results. There has been a lack of numerical studies mixing, leaving, secondary kinetic energy, and clearance losses,
on the effect of the Reynolds number on centrifugal compressor which do not depend on the Reynolds number.
performance and losses. Because the loss generation in low-Reynolds-number com-
Several correction equations for efficiency are published in pressors is not well known yet, the efficiency correction equa-
literature. Some of the correction equations are represented in tions should be used carefully. It should be noted that the
Tab. 1. In the past, the old empirical formula was used to cor- correction equation presented in the standard ASME PTC 10
rect the compressor efficiency. However, its weaknesses are the 1997 Performance Test Code on Compressors and Exhausters [6]
neglect of the surface roughness effect and the assumption of the does not divide the losses into Reynolds-number-dependent and
constant fraction of the losses independent of the Reynolds num- Reynolds-number-independent losses and the correction equa-
ber [1]. tion in the standard ISO 5389:2005 Turbocompressors - Perfor-
In 1985, Casey [2] introduced a correction equation where a mance test code [7] is just a modification of the old empirical
change in the friction factor of the equivalent pipe flow due to a formula.
change in the Reynolds number causes a change in the stage ef- The literature review shows that the previous studies of the
ficiency. The correction equation of Casey [2] has recently been effect of the Reynolds number on centrifugal compressor perfor-
modified [1, 3]. mance have been experimental and efficiency correction equa-
Heß and Pelz [4] published an efficiency correction equation tions are derived based on those empirical results. There is a
for axial turbomachines, which is based on an old empirical for- lack of investigations where the effect of the Reynolds number
mula. The change in efficiency due to the Reynolds number is on centrifugal compressor performance and losses is studied nu-
related to the Reynolds-dependent loss fraction, which is a func- merically. The aim of this paper is to investigate how well the
tion of the flow coefficient. numerical results follow the correction equations found in the
In 2013, Pelz and Stonjek [5] published their version of literature. The loss generation in the impeller blade passages is
the correction equation, which does not include empirical coeffi- also studied in order to find out which loss generation mecha-
cients. The friction coefficient is that of an equivalent pipe flow nism has the most potential to be reduced or eliminated in very

2 Copyright © 2016 by ASME

Downloaded From: https://proceedings.asmedigitalcollection.asme.org on 06/30/2019 Terms of Use: http://www.asme.org/about-asme/terms-of-use


TABLE 1. SUMMARY OF THE EFFICIENCY CORRECTION ψ, and impeller tip speed Mach number MaU ) were kept con-
EQUATIONS REPRESENTED IN LITERATURE. stant, except the Reynolds number, which decreased as the com-
pressor was downscaled.
Reference Equation According to Japikse [17], from the constant impeller tip
1−η
h
Reref
in speed Mach number MaU , the relation for the rotational speed of
Old empirical formula [10] = a + (1 − a) (1)
1−ηref Re the downscaled compressor can be written as:
Casey (1985) [2] ∆η = − µc0 ∆ f (2)
 n
1−η Reref
Heß & Pelz (2010) [4] 1−ηref = (1 − b) + b Re (3) 1
Nscaled = Nbaseline . (7)
Casey & Robinson (2011) [1] ∆η = − Bfref
ref
∆f (4) SF
Bref
Dietmann & Casey (2013) [3] ∆η = − fref ∆ f (5)
Similarly from the constant flow coefficient φ , the relation for
Pelz & Stonjek (2013) [5] ∆η = − 1−η ref
cf,ref ∆cf (6)
the mass flow rate of the downscaled compressor can be written
as:

small centrifugal compressors.


In this paper, the effect of the Reynolds number on centrifu- qm,scaled = SF 2 qm,baseline . (8)
gal compressor performance and losses is studied numerically.
An industrial, high-speed compressor and a test case compres-
sor Radiver [8, 9] are downscaled. The downscaled compressors Ernst et al. [18] recommended the use of the diameter scal-
and the original ones are modeled in the chord Reynolds num- ing factor in the range of 0.8 to 1.2 in order to keep the effect
ber range of 0.8 · 105 ...17 · 105 . The numerical results are vali- of the Reynolds number small. In this paper, the scaling fac-
dated against experimental data and compared with the correc- tors in the range of 0.05 to 0.9 are used in order to study the
tion equations found in the literature. Reynolds number effect. Compressor 1 is downscaled 10 times
with scaling factors varying from 0.05 to 0.9, whereas Radiver
is downscaled 6 times with scaling factors varying from 0.05 to
0.8. The chord Reynolds number is varied within the range from
STUDIED COMPRESSORS
0.8 · 105 to 17 · 105 .
The first studied compressor is an unshrouded high-speed
centrifugal compressor with a design pressure ratio of 1.78 and
specific speed of 0.8. The impeller has seven full and seven
splitter blades with a 40◦ backsweep. The relative tip clearance, NUMERICAL METHODS
t2 /b2 , is 0.052. The same compressor has been widely tested and To investigate the efficiency of the downscaled compressors,
modeled at Lappeenranta University of Technology [11–16]. In they were modeled by the commercial CFD software ANSYS
the following, this compressor is referred to as Compressor 1. CFX in steady state. The computational domains of Compressor
The second studied compressor is a widely known test case 1 and Radiver consisted of an inlet section, impeller, vaneless
compressor called Radiver in the Institute of Jet Propulsion and diffuser with parallel walls, and an exit section. The exit section
Turbomachinery of RWTH Aachen with a specific speed of 0.69 between the diffuser outlet and computational domain outlet was
[8,9]. The impeller has 15 full blades with a 38◦ backsweep. The used to eliminate the reversed flow at the diffuser outlet due to
relative tip clearance is 0.045. Both Compressor 1 and Radiver an outlet boundary. The computational domains of Compressor
are studied with vaneless diffusers. 1 and Radiver are shown in Fig. 1.
The test case compressor Radiver is used only as a compar- The SST k − ω model was used to model turbulence. The
ison to study the performance deterioration of the downscaled values of the non-dimensional wall distance were below unity
compressors. The loss generation in the impeller blade passages on most of the surfaces. The frozen rotor approach was used to
is studied in more detail in Compressor 1. model the interaction between rotating and stationary parts. For
In order to study the effect of the Reynolds number on every case, the total pressure and temperature were given at the
the compressor performance, the compressors were downscaled. inlet boundary of the computational domain and mass flow rate
All geometric dimensions of the compressors were downscaled at the outlet boundary. Second order discretization was used for
with the same scaling factor as the impeller outlet diameter, the equations. The converged results were achieved when there
D
SF = D 2,scaled . Also the same gas properties were used for the was no change in residuals, efficiency, or mass flow rate at the
2,baseline
downscaled compressors as for the baseline compressor. All the inlet. The values of residuals were of the order of 10−4 at the
dimensionless numbers (flow coefficient φ , pressure coefficient convergence.

3 Copyright © 2016 by ASME

Downloaded From: https://proceedings.asmedigitalcollection.asme.org on 06/30/2019 Terms of Use: http://www.asme.org/about-asme/terms-of-use


1.28
Measurement Compressor 1
1.24
CFD Compressor 1
1.20 Measurement Radiver
1.16 CFD Radiver
1.12

ηtt /ηtt,ref
1.08
1.04
1.00
0.96
0.92

0.8 1.0 1.2 1.4


qm/qm,des

FIGURE 1. THE 3D COMPUTATIONAL DOMAINS A) COM- FIGURE 3. VALIDATION OF COMPUTATIONAL RESULTS OF


PRESSOR 1, B) RADIVER (NOT IN SCALE). EFFICIENCY AGAINST EXPERIMENTAL DATA.

1.02 1.02 1.28


Measurement Compressor 1
1.24
πtt /πtt,ref
ηtt /ηtt,ref

CFD Compressor 1
1.00 1.00 1.20 Measurement Radiver
1.16 CFD Radiver

πtt /πtt,ref
0.98 0.98 1.12
1 2 3 4 1.08
1.02 1.02 1.04
1.00 1.00
πtt /πtt,ref
ηtt /ηtt,ref

1.00
0.98 0.98 0.96
0.96 Efficiency 0.96
0.94 Pressure ratio 0.94 0.92
0.92 0.92
1 2 3 4 0.8 1.0 1.2 1.4
6 qm/qm,des
Number of cells, 10

FIGURE 2. MESH INDEPENDENCE OF COMPRESSOR 1 (UP- FIGURE 4. VALIDATION OF COMPUTATIONAL RESULTS OF


PER) AND RADIVER (LOWER). PRESSURE RATIO AGAINST EXPERIMENTAL DATA.

Validation Against Experimental Data


Mesh Independence Study The computational models were also validated against ex-
For the mesh independence study, three meshes with 0.8, perimental results in the compressors with original sizes without
1.9, and 4.3 million computational cells were used for Com- scaling. The computational and measured results of the total-
pressor 1 and three meshes with 0.7, 1.7, and 3.8 million cells to-total efficiency and pressure ratio are shown as functions of
for Radiver. As a result of the mesh independence study, the the non-dimensionalized mass flow rate for Compressor 1 and
meshes with 1.9 and 1.7 million cells were chosen for Compres- Radiver in Figs. 3 and 4, respectively. The efficiency and pres-
sor 1 and Radiver, respectively. The studied variables in the mesh sure ratio are normalized by the measured value at the design
independence study were the total-to-total efficiency and total- point and the mass flow rate is non-dimensionalized by the mass
to-total pressure ratio between the computational domain inlet flow rate at the design point. For computational results, the dis-
and diffuser outlet. The discretization error was estimated using cretization error is presented but the sizes of the error bars are in
the procedure presented by Celik et al. [19]. The estimated dis- the order of the line width.
cretization error is shown in Fig. 2 , which presents the results of The validation results show that the computational model
the mesh independence study for Compressor 1 (upper subplot) over-predicts the efficiency and pressure ratio in both cases, but
and Radiver (lower subplot). The meshes of the baseline com- captures the trend. It must be noted that the computational ef-
pressors are scaled for the downscaled compressors, i.e. having ficiency and pressure ratio are calculated between the compu-
the same number of cells in both baseline and downscaled cases. tational domain inlet and diffuser outlet, whereas the measure-

4 Copyright © 2016 by ASME

Downloaded From: https://proceedings.asmedigitalcollection.asme.org on 06/30/2019 Terms of Use: http://www.asme.org/about-asme/terms-of-use


1.1 1.1
SF = 1.0 SF = 1.0

1.0 SF = 0.05 1.0 SF = 0.05


η/ηref

η/ηref
3.6% 4.6%
0.9 Casey (1985), Reb2 0.9 Casey (1985), Reb2
Dietmann & Casey (2013), Rec Dietmann & Casey (2013), Rec
Pelz & Stonjek (2013), Reb2 Pelz & Stonjek (2013), Reb2
Recrit/Reref Recrit/Reref
Compressor 1 Radiver
0.8 −1 0
0.8 −1 0
10 10 10 10
Re/Reref Re/Reref

FIGURE 5. VARIATION OF EFFICIENCY η IN THE CASE OF FIGURE 6. VARIATION OF EFFICIENCY η IN THE CASE OF
DOWNSCALED COMPRESSOR 1. SF REFERS TO SCALING FAC- DOWNSCALED RADIVER. SF REFERS TO SCALING FACTOR.
TOR.

downscaled Radiver, the tip clearance is also extremely small (25


ments are done between the compressor inlet and outlet for both µm). Thus, it might be that the additional losses accounted by
Compressor 1 and Radiver. Therefore, the pressure loss in the vo- the correction equation are generated by the relatively increased
lute is neglected in the computational results and can be seen as blade thickness and other geometrical changes, which are nec-
a part of the difference between the computational and measured essary due to the manufacturing limitations but which are not
values. The losses due to disk friction are not either captured by accounted for in this study. Therefore, in the following sections
the computational model. The part of the difference between the the downscaled compressor SF = 0.1 is studied in detail instead
computational and measured results is also due to the weakness of the smallest one SF = 0.05, which is very difficult to manu-
of two-equation models in predicting all the losses. facture with the available techniques.

RESULTS Blockage in the Blade Passage


Performance Deterioration of the Downscaled Com- It is generally known that the losses in low-Reynolds-
pressors number compressors are caused by the relatively larger boundary
Figures 5 and 6 show how the efficiency decreases due to layer thickness, surface roughness, blade thickness, and tip clear-
a decreasing Reynolds number in the downscaled Compressor 1 ance. In this study, any increase in the relative size of the other
and Radiver, respectively. Only three correction equations are variables except the boundary layer thickness is eliminated by
plotted in Figs. 5 and 6 for the sake of clarity. The selected the scaling all geometric dimensions of the compressors with the
correction equations are the best ones according to the literature same scaling factor. Therefore, only losses due to the increased
review and they also give the best fit for the computational re- boundary layer thickness are investigated.
sults. The following figures (7 and 8) show the normalized merid-
It is clear that the efficiency of the downscaled compressors ional velocity profiles in pitchwise (blade-to-blade) direction at
Compressor 1 and Radiver follows quite precisely the correc- two different meridional locations at mid-span for the baseline
tion equation published by Dietmann and Casey [3]. However, compressor (SF = 1.0) and two downscaled ones (SF = 0.1 and
the prediction is not as precise well below (SF = 0.05) the crit- 0.2). The results are shown only for these three cases because
ical chord Reynolds number (200 000). The difference between there is no significant change in the results between SF = 1.0
the correction equation and the numerical results (SF = 0.05) is and SF = 0.2. The most significant changes happen when the
3.6% for Compressor 1 and 4.6% for Radiver. chord Reynolds number is decreased below the critical value of
It should be noted that the rotational speed of the smallest 200 000 [1]. There is no significant difference in velocity pro-
compressor (SF = 0.05) is extremely high for both Compres- files between the planes 60% (not shown) and 80% (Fig. 7) from
sor 1 and Radiver (360 krpm and 563 krpm, respectively). Also the impeller inlet, except the weak low meridional velocity re-
the blade thickness is significantly smaller than the lower limit gion on the full blade suction side in the case of the downscaled
(0.35 mm) presented by Came and Robinson [20]. In the smallest compressor SF = 0.1. Near the trailing edge (Fig. 8), the region

5 Copyright © 2016 by ASME

Downloaded From: https://proceedings.asmedigitalcollection.asme.org on 06/30/2019 Terms of Use: http://www.asme.org/about-asme/terms-of-use


1.0 1.0
SF = 0.1 SF = 0.1
SB SF = 0.2 FB SB SF = 0.2 FB
0.8 SF = 1.0 0.8 SF = 1.0

SS PS SS PS SS PS SS PS
cm/U2 [−]

cm/U2 [−]
0.6 0.6

0.4 0.4

0.2 0.2

0.0 0.0
0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0
Dimensionless pitchwise direction [−] Dimensionless pitchwise direction [−]

FIGURE 7. NORMALIZED MERIDIONAL VELOCITY COMPO- FIGURE 8. NORMALIZED MERIDIONAL VELOCITY COMPO-
NENT AS A FUNCTION OF THE DIMENSIONLESS PITCHWISE NENT AS A FUNCTION OF THE DIMENSIONLESS PITCHWISE
DIRECTION AT THE PLANE 80% FROM THE IMPELLER INLET DIRECTION NEAR THE IMPELLER TRAILING EDGE AND 50%
IN THE MERIDIONAL DIRECTION AND 50% FROM THE HUB IN FROM THE HUB IN THE SPANWISE DIRECTION.
THE SPANWISE DIRECTION.
1.0
25%
SB 50% FB
of low meridional velocity is visible in every case, but it is the 0.8 75%
strongest in the downscaled compressor SF = 0.1.
The results shown in Figs. 7 and 8 indicate that the merid- SS PS SS PS
cm/U2 [−] 0.6
ional velocity increases in the blade passages with a decreasing
Reynolds number. This phenomenon results from an increased 0.4
blockage due to thicker boundary layers. Boundary layer de-
velopment is studied in more detail later in Fig. 10, but first, the 0.2
differences in velocity profiles at different spanwise locations are
examined. 0.0
0.0 0.2 0.4 0.6 0.8 1.0
The normalized meridional velocity component is shown at Dimensionless pitchwise direction [−]
the plane 80% from the impeller inlet in the meridional direc-
tion at three different spanwise locations (25%, 50%, and 75%
from the hub) for the scaled compressor SF = 0.1 in Fig. 9. The FIGURE 9. NORMALIZED MERIDIONAL VELOCITY COMPO-
results show that there is no significant difference in the veloc- NENT AS A FUNCTION OF THE DIMENSIONLESS PITCHWISE
ity profile in the spanwise direction except near the shroud (75% DIRECTION AT THE PLANE 80% FROM THE IMPELLER IN-
from the hub) where the lower values of meridional velocity are LET IN THE MERIDIONAL DIRECTION FOR THE SCALED COM-
visible. Hence, boundary layer development in the baseline com- PRESSOR SF = 0.1.
pressor (SF = 1.0) and two downscaled ones (SF = 0.1 and 0.2)
is studied only at the meridional plane 80% from the impeller in- span the increase in boundary layer thickness due to a decreased
let and at mid-span. The normalized meridional velocity profiles Reynolds number is in the same order as at the plane 80% from
are shown more closely for the full blade suction side in Fig. 10. the impeller inlet.
From Fig. 10, one can see that the boundary layer thickens Then, the boundary layer thickness in the spanwise direction
with a decreasing compressor size. The thickening of the bound- is investigated. Figure 11 shows the normalized meridional ve-
ary layer is similarly visible on the suction side of the splitter locity from hub to shroud near the impeller trailing edge in the
blade and on the pressure side of the full blade. There is no middle of the blade passage between the full blade suction side
visible increase in the boundary layer thickness on the splitter and splitter blade pressure side. There is no significant differ-
blade pressure side. In Fig. 10, the ”free stream velocity” cm,∞ in ence in the boundary layer thickness between the baseline and
the boundary layer definition is defined as the maximum velocity downscaled cases, but it is obvious that the meridional velocity
near the blade surface, i.e. cm,∞ /U2 ≈ 0.2. is increased in the blade passage at mid-span due to the larger
Velocity profiles were also investigated in detail close to the region of the low meridional velocity.
trailing edge (similar to Fig. 10). The results show that at mid- A larger region of low meridional velocity (approx. 10%)

6 Copyright © 2016 by ASME

Downloaded From: https://proceedings.asmedigitalcollection.asme.org on 06/30/2019 Terms of Use: http://www.asme.org/about-asme/terms-of-use


0.75 Boundary layer
thickness
cm = 0.99cm,∞

0.50
cm /U2 [−]

0.25 SF = 0.1
SF = 0.2
SF = 1.0 FIGURE 12. MERIDIONAL VELOCITY CONTOURS AT THE
0.00
0 2 4 IMPELLER TRAILING EDGE FOR THE SCALING FACTORS 0.1
Pitch [%] AND 1.0. THE REGION OF THE LOW MERIDIONAL VELOCITY
IS MARKED BY A CIRCLE.
FIGURE 10. NORMALIZED MERIDIONAL VELOCITY COM-
PONENT AS A FUNCTION OF THE DIMENSIONLESS PITCH-
Blade Loading of Compressor 1
WISE DIRECTION AT THE PLANE 80% FROM THE IMPELLER
Blade loading distributions along the full and splitter blade
INLET IN THE MERIDIONAL DIRECTION AND 50% FROM THE
pressure and suction sides are shown in Fig. 13 at the tip. The
HUB IN THE SPANWISE DIRECTION NEAR THE FULL BLADE
distributions at the mid-span and hub are similar. The blade
(FB) SUCTION SIDE (SS).
loading distribution of a full blade is shown by black lines and
that of a splitter blade by red lines. To calculate the blade load-
ing, the static pressure on the blade surface is normalized by the
1.0 static pressure at the impeller inlet. As shown in the previous
SF = 0.1
SF = 0.2
sections, the performance of the downscaled compressors is de-
0.8 creased. The performance deterioration can also be seen as lower
spanwise direction [-]

SF = 1.0
blade loading. At first, the results indicate that the blade loading
0.6 is equally deteriorated along the blade surface from the leading
Dimensionless

edge to the trailing edge. However, it is clear that the blade load-
ing is deteriorated relatively more in the case of SF = 0.1 on the
0.4
full blade suction side (black dotted line) near the trailing edge
(subplot on the lower right corner) and at the tip on the splitter
0.2 blade pressure side (red dotted line in the subplot on the lower
left corner), especially in the region from the leading edge to the
0.0 middle of the splitter blade length.
−0.2 0.0 0.2 0.4 0.6
cm /U2 [−] Previously it was observed that the increase in the boundary
layer thickness on the splitter blade pressure side is not visible
at the mid-span. However, the blade loading is deteriorated rel-
FIGURE 11. NORMALIZED MERIDIONAL VELOCITY COM-
atively more on the splitter blade pressure side especially near
PONENT FROM HUB (0) TO SHROUD (1) NEAR THE IMPELLER
the splitter blade leading edge at the tip due to the tip leakage
TRAILING EDGE AND IN THE MIDDLE OF THE BLADE PAS-
flow. The mechanism behind this observation is that the part of
SAGE BETWEEN THE FULL BLADE SUCTION SIDE AND SPLIT-
the tip leakage flow over the full blade migrates with the tip leak-
TER BLADE PRESSURE SIDE.
age flow over the splitter blade [15], resulting in decreased blade
loading.

can be seen in Fig. 12, where meridional velocity contours are Shear Stress in Compressor 1
shown at the impeller trailing edge for the baseline and down- The results in previous sections showed that the boundary
scaled compressor. In the downscaled compressor (SF = 0.1), layers in the pitchwise or spanwise direction are not significantly
the size of the low meridional velocity region is increased in the increased near the trailing edge where there is a significant dete-
spanwise direction, resulting in increased meridional velocity in rioration in the blade loading especially on the full blade suction
the middle of the blade passage. side. Therefore, it is concluded that the increased blockage is not

7 Copyright © 2016 by ASME

Downloaded From: https://proceedings.asmedigitalcollection.asme.org on 06/30/2019 Terms of Use: http://www.asme.org/about-asme/terms-of-use


1.6 1.0

Normalized shear strain rate [−]


1.4 SF = 0.1
ps/ps,1 [−]

SB FB
1.2 SF = 0.2
SF = 0.1, FB 0.8 SF = 1.0
1.0 SF = 1.0, FB
SF = 0.1, SB
0.8
SF = 1.0, SB
SS PS SS PS
0.6
0.6
0.0 0.2 0.4 0.6 0.8 1.0
Dimensionless streamwise direction [−]
1.4 0.4
1.6
1.2
0.2
1.0 1.4

0.8 0.0
1.2 0.0 0.2 0.4 0.6 0.8 1.0
0.6 Dimensionless pitchwise direction [−]
0.2 0.4 0.6 0.8 1.0

FIGURE 14. NORMALIZED SHEAR STRAIN RATE AS A FUNC-


FIGURE 13. BLADE LOADING AT THE TIP FOR THE FULL
TION OF THE DIMENSIONLESS PITCHWISE DIRECTION NEAR
BLADE (BLACK) AND SPLITTER BLADE (RED) ON THE PRES-
THE IMPELLER TRAILING EDGE AND 75% FROM THE HUB IN
SURE AND SUCTION SIDES.
THE SPANWISE DIRECTION.

the only reason for the decreased blade loading and compressor 1.0

Normalized shear strain rate [−]


SF = 0.1
performance. SB FB
SF = 0.2
The normalized shear strain rate is shown in pitchwise 0.8 SF = 1.0
(blade-to-blade) direction near the trailing edge at two different SS PS SS PS
spanwise locations in Figs. 14 and 15. The black bars show the 0.6
locations of the full and splitter blades. Here, the shear strain
rate is one of thescalar invariants of the rate-of-strain tensor 0.4
∂u
Sij = 12 ∂∂ uxi + ∂ xj and it is defined as S = 12 Sij Sij [21]. In Figs.
j i
14 and 15, the maximum value of the shear strain rate in SF = 0.1 0.2
at the trailing edge is used to normalize the data.
The results indicate that the shear stress increases signifi- 0.0
0.0 0.2 0.4 0.6 0.8 1.0
cantly from 75% to 98% span and it is the largest close to the Dimensionless pitchwise direction [−]
stationary shroud (Fig. 15). From the hub to 75% span there is
no significant difference in the strain rate. One can notice also FIGURE 15. NORMALIZED SHEAR STRAIN RATE AS A FUNC-
that the strain rate has peak values on the blade surfaces because TION OF THE DIMENSIONLESS PITCHWISE DIRECTION NEAR
of the boundary layers. The shear stress is significantly stronger THE IMPELLER TRAILING EDGE AND 98% FROM THE HUB IN
near the shroud in the downscaled compressor than in the base- THE SPANWISE DIRECTION.
line one. Figure 16 shows that also vorticity (∇ ×~u) increases
from the baseline compressor to the downscaled one near the
walls (hub, shroud, and blade surfaces). Increased shear stress
and boundary layer thickness produce more losses, which is rec-
ognized as an increase in entropy (Fig. 17).
The scalar invariant of the rate-of-strain tensor S refers to
high dissipation rate [21]. Therefore, the following conclusion
is drawn from Figs. 14, 15, 16, and 17: the regions of the high
vorticity (Fig. 16) correlate with the regions of the high dissipa-
tion rate (shear strain rate, Figs. 14 and 15) resulting in increased
losses (entropy, Fig. 17). One reason for the increased vorticity
in the downscaled compressors might be the increased centrifu-
gal force tending to separate the high and low momentum fluids.
FIGURE 16. VORTICITY CONTOURS IN THE IMPELLER FOR
Figure 18 shows how the increase in entropy is divided be-
THE SCALING FACTORS 0.1 AND 1.0.
tween different causes of loss. The values of entropy are inves-

8 Copyright © 2016 by ASME

Downloaded From: https://proceedings.asmedigitalcollection.asme.org on 06/30/2019 Terms of Use: http://www.asme.org/about-asme/terms-of-use


CONCLUSIONS
The performance deterioration of low-Reynolds-number
compressors was studied numerically in this paper. Two cen-
trifugal compressors, one with full and splitter blades and an-
other with only full blades, were modeled and downscaled using
scaling factors from 0.05 to 0.9 in order to study the effect of the
Reynolds number in the range of chord Reynolds numbers from
0.8 · 105 to 17 · 105 . The computational results were validated
against experimental data and compared with the efficiency cor-
rection equations found in the literature. Furthermore, a detailed
analysis of the flow field was performed for one of the compres-
FIGURE 17. SPECIFIC ENTROPY CONTOURS IN THE IM- sors.
PELLER FOR THE SCALING FACTORS 0.1 AND 1.0. According to the results, with the decreasing Reynolds num-
ber the compressor performance decreases quite precisely as the
efficiency correction equation published by Dietmann and Casey
100 [3] predicts. However, a difference between the numerical re-
sults and correction equation was observed. The difference was
Increase in entropy [%]

75 3.6% for Compressor 1 and 4.6% for Radiver. It might be that


this difference below the critical chord Reynolds number (200
50
000) is due to the manufacturing tolerances which are neglected
in this numerical study but accounted for in the correction equa-
tion based on the experimental data. This assumption should be
25
further studied.
The results show that there is no significant difference in
0
compressor performance or flow fields when the Reynolds num-
al
S)

es
S)

b
e

hu
nc

t
To
BS
BS

ad

ber is above the critical value (200 000). The most significant
ra

bl
(S
(F

th
ea

e
e

changes happen when the chord Reynolds number is decreased


e

at
cl

th
ak

ak
p

L.

at
W
Ti

B.

below the critical value. The significant changes are observed in


L.
B.

the boundary layer thickness, wake size, and shear stress near the
shroud. Also the decreased blade loading indicated lower perfor-
FIGURE 18. INCREASE IN SPECIFIC ENTROPY BETWEEN
mance with the decreasing Reynolds number.
CASES SF = 1.0 AND SF = 0.1. B.L. REFERS TO THE BOUNDARY
The results indicated that the increased shear stress due to
LAYER.
larger velocity differences resulting from high vorticity possi-
bly because of high rotational speeds is as significant loss pro-
duction mechanism in low-Reynolds-number compressors as the
increased boundary layer thickness on the blade surfaces. The
tigated at the impeller trailing edge in the cases SF = 1.0 and large increase in entropy due to the increased shear stress was
SF = 0.1. Tip clearance refers to the rectangular plane in the observed near the shroud where also the increased values of vor-
tip clearance. The wakes on the full (FBSS) and splitter (SBSS) ticity were observed. However, it would be interesting to study
blade suction sides are defined as low meridional velocity regions loss production mechanisms if the size of the compressor is not
(cm /U2 ≤ 0.2) like in the study of Eckardt [22]. The boundary downscaled but the Reynolds number is decreased by changing
layer edges near the hub and blades are defined as cm = 0.99cm,∞ , the inlet conditions. The blockage might be a more significant
where cm,∞ /U2 ≈ 0.2. This is just a rough estimation of the loss loss production mechanism than shear stress resulting from the
fractions. increased vorticity if the inlet conditions were changed instead
Figure 18 clearly shows that the major part of the entropy of the compressor size and rotational speed.
increase is located in the tip clearance region and in the bound- In the future, the results of Radiver will be published and it
ary layers near the blade surfaces, where the shear stress is sig- would be important to validate CFD codes against experimental
nificantly stronger due to high vorticity. The study of mass-flow- data at low Reynolds numbers, especially under the critical chord
averaged specific entropy at the observation planes in the diffuser Reynolds number 200 000. Unfortunately, the detailed analysis
showed that the wakes were mixed out at the location R = 1.1R2 . of the flow field inside an impeller of the micro-scale centrifugal
All the findings above emphasize the important contribution of compressor is still very hard with the available measuring tech-
compressor downscaling on the loss share. niques.

9 Copyright © 2016 by ASME

Downloaded From: https://proceedings.asmedigitalcollection.asme.org on 06/30/2019 Terms of Use: http://www.asme.org/about-asme/terms-of-use


ACKNOWLEDGMENT and Backman, J., 2012. “Experimental Study of the Effect
This research is part of the project funded by the Academy of the Tip Clearance to the Diffuser Flow Field and Stage
of Finland under grant number 274897. Performance of a Centrifugal Compressor”. In Proceedings
of the ASME Turbo Expo 2012. Paper No. GT2012-68445
p. 8. June 11-15, 2012, Copenhagen, Denmark.
REFERENCES [13] Jaatinen-Värri, A., Röyttä, P., Turunen-Saaresti, T., and
[1] Casey, M., and Robinson, C., 2011. “A Unified Correc- Grönman, A., 2013. “Experimental Study of Centrifugal
tion Method for Reynolds Number, Size, and Roughness Compressor Vaneless Diffuser Width”. Journal of Mechan-
Effects on the Performance of Compressors”. Proceedings ical Science and Technology, 27(4), pp. 1011–1020.
of the Institution of Mechanical Engineers, Part A: Journal [14] Jaatinen-Värri, A., Turunen-Saaresti, T., Röyttä, P.,
of Power and Energy, 225(7), pp. 864–876. Grönman, A., and Backman, J., 2013. “Experimental Study
[2] Casey, M., 1985. “The Effects of Reynolds Number on the of Centrifugal Compressor Tip Clearance and Vaneless Dif-
Efficiency of Centrifugal Compressor Stages”. Journal of fuser Flow Fields”. Proceedings of the Institution of Me-
Engineering for Gas Turbines and Power, 107, pp. 541– chanical Engineers, Part A: Journal of Power and Energy,
548. 227(8), pp. 885–895.
[3] Dietmann, F., and Casey, M., 2013. “The Effects of [15] Jaatinen-Värri, A., Grönman, A., Turunen-Saaresti, T., and
Reynolds Number and Roughness on Compressor Perfor- Backman, J., 2014. “Investigation of the Stage Performance
mance”. In Proceedings of the 10th European Conference and Flow Fields in a Centrifugal Compressor with a Vane-
on Turbomachinery: Fluid Dynamics and Thermodynam- less Diffuser”. International Journal of Rotating Machin-
ics, J. Backman, G. Bois, and O. Leonard, eds., pp. 532– ery, 2014, pp. Article ID 139153, 10 pages.
542. April 15-19, 2013, Lappeenranta, Finland. [16] Jaatinen-Värri, A., Turunen-Saaresti, T., Grönman, A.,
[4] Heß, M., and Pelz, P., 2010. “On Reliable Performance Backman, J., and Tiainen, J., 2015. “Numerical Investi-
Prediction Of Axial Turbomachines”. In Proceedings of gation of Centrifugal Compressor Tip Clearance”. In Pro-
ASME Turbo Expo 2010: Power for Land, Sea and Air, ceedings of the ASME Turbo Expo 2015: Turbine Techni-
Vol. 7, pp. 139–149. Paper No. GT2010-22290. June 14- cal Conference and Exposition. Paper No. GT2015-43199,
18, 2010, Glasgow, UK. p.10. June 15-19, 2015, Montreal, Canada.
[5] Pelz, P., and Stonjek, S., 2013. “The Influence of Reynolds [17] Japikse, D., 1996. Centrifugal Compressor Design and Per-
Number and Roughness on the Efficiency of Axial and Cen- formance. Wilder (VT): Concepts ETI. p. appr. 500. ISBN
trifugal Fans - a Physically Based Scaling Method”. Jour- 0-933283-03-2.
nal of Engineering for Gas Turbines and Power, 135(5), [18] Ernst, B., Kammeyer, J., and Seume, J., 2011. “Improved
pp. Article ID 052601, 8 pages. Map Scaling Methods for Small Turbocharger Compres-
[6] ASME PTC 10, 1997. Performance Test Code on Com- sors”. In Proceedings of ASME Turbo Expo 2011, Vol. 3,
pressors and Exhausters. p. 188. pp. 733–744. Paper No. GT2011-45345. June 6-10, 2011,
[7] ISO 5389, 2005. Turbocompressors - Performance test Vancouver, British Columbia, Canada.
code. p. 142. [19] Celik, I., Ghia, U., Roache, P., Freitas, C., Coleman, H.,
[8] Ziegler, K., Gallus, H., and Niehuis, R., 2003. “A Study on and Raad, P., 2008. “Procedure for Estimation and Report-
Impeller-Diffuser Interaction - Part I: Influence on the Per- ing of Uncertainty Due to Discretization in CFD Applica-
formance”. Journal of Turbomachinery, 125(1), pp. 173– tions”. Journal of Fluids Engineering, 130(7), pp. Article
182. ID 078001, 4 pages.
[9] Ziegler, K., Gallus, H., and Niehuis, R., 2003. “A Study on [20] Came, P. M., and Robinson, C. J., 1999. “Centrifugal Com-
Impeller-Diffuser Interaction - Part II: Detailed Flow Anal- pressor Design”. Proceedings of the Institution of Mechan-
ysis”. Journal of Turbomachinery, 125(1), pp. 183–192. ical Engineers, Part C: Journal of Mechanical Engineering
[10] Wiesner, F. J., 1979. “A New Appraisal of Reynolds Num- Science, 213(2), pp. 139–155.
ber Effects on Centrifugal Compressor Performance”. Jour- [21] Soria, J., Sondergaard, R., Cantwell, B., Chong, M., and
nal of Engineering for Gas Turbines and Power, 101(3), Perry, A., 1994. “A Study of the Fine-Scale Motions of
pp. 384–392. Incompressible Time-Developing Mixing Layers”. Physics
[11] Jaatinen, A., Grönman, A., Turunen-Saaresti, T., and of Fluids, 6(2), pp. 871–884.
Röyttä, P., 2011. “Effect of Vaneless Diffuser Width on the [22] Eckardt, D., 1976. “Detailed Flow Investigations Within a
Overall Performance of a Centrifugal Compressor”. Pro- High-Speed Centrifugal Compressor Impeller”. Journal of
ceedings of the Institution of Mechanical Engineers, Part Fluids Engineering, 98(3), pp. 390–399.
A: Journal of Power and Energy, 225(5), pp. 665–673.
[12] Jaatinen, A., Turunen-Saaresti, T., Grönman, A., Röyttä, P.,

10 Copyright © 2016 by ASME

Downloaded From: https://proceedings.asmedigitalcollection.asme.org on 06/30/2019 Terms of Use: http://www.asme.org/about-asme/terms-of-use

You might also like