J Fluids Engineering 2009 Vol 131 N5
J Fluids Engineering 2009 Vol 131 N5
J Fluids Engineering 2009 Vol 131 N5
Fluids Engineering
Published Monthly by ASME
RESEARCH PAPERS
Flows in Complex Systems
051101
051102
051103
051104
051105
051106
051202
PUBLISHING STAFF
Managing Director, Publishing
P. DI VIETRO
Manager, Journals
C. MCATEER
Production Coordinator
A. HEWITT
Multiphase Flows
051301
051302
051401
051402
051403
Downloaded 03 Jun 2010 to 171.66.16.159. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Contents continued
Journal of Fluids Engineering
MAY 2009
TECHNICAL BRIEFS
054501
Downloaded 03 Jun 2010 to 171.66.16.159. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Many centrifugal pumps have a suction velocity profile, which is nonuniform, either by
design like in double-suction pumps, sump pumps, and in-line pumps, or as a result of an
installation close to an upstream disturbance like a pipe bend. This paper presents an
experimental study on the effect of a nonuniform suction velocity profile on performance
of a mixed-flow pump and hydrodynamic forces on the impeller. In the experiments, a
newly designed dynamometer is used, equipped with six full Wheatstone bridges of strain
gauges to measure the six generalized force components. It is placed in between the shaft
of the pump and the impeller and corotates with the rotor system. A high accuracy is
obtained due to the orthogonality of bridge positioning and the signal conditioning electronics embedded within the dynamometer. The suction flow distribution to the pump is
adapted using a pipe bundle situated in the suction pipe. Results of measurements show
the influence of the suction flow profile and blade interaction on pump performance and
forces. Among the most important observations are a backward whirling motion of the
rotor system and a considerable steady radial force. DOI: 10.1115/1.3089539
Keywords: fluid-induced forces, hydrodynamics, mixed-flow pump, nonuniform suction
flow
Introduction
Forces on pump shafts may have several causes. The origin can
be mechanical, such as mass unbalance or misalignment, or induced by the working fluid. Generally, a nonuniform pressure distribution at the periphery of the impeller is regarded as a cause for
a steady lateral force, like the one encountered in pumps equipped
with a volute type of casing. Hydrodynamic excitation forces result from a variety of unsteady flow phenomena such as stall, flow
recirculation, cavitation, and blade interaction, while hydrodynamic reaction forces on the other hand may result from rotor
whirl or precessing motion relative to the casing.1.
Agostinelli et al. 2 and Iversen et al. 3 were the first to
measure the steady lateral force on an impeller in a vaneless volute. Domm and Hergt 4 did similar experiments in which they
demonstrated that impeller forces are dependent on rotor eccentricity. Hergt and Krieger 5 extended these measurements for an
impeller, which was placed eccentrically in a vaned diffuser.
Uchida et al. 6 measured the unsteady lateral forces on an impeller in a volute type of casing. In all of these experiments, force
sensing devices of different designs were used to measure reaction
forces in the bearing support structures. Chamieh et al. 7 used an
external balance to measure the hydrodynamic stiffness matrix for
a centrifugal impeller in a volute, whirling at a very low speed.
Internal balances, or so-called dynamometers, were used by
subsequent investigations to assess the fluid-induced rotordynamic
behavior of centrifugal pumps. These force sensing devices are
located between the impeller and the shaft so as to directly measure forces acting on the impeller. Jerry et al. 8 constructed a
dynamometer with four square posts parallel to the shaft, onto
which were mounted nine full Wheatstone bridges of strain
gauges to measure six generalized force components. Rotordynamic forces were measured for a radial-type centrifugal impeller
in a vaneless volute forced into a whirling motion. Further re-
2
Contributed by the Fluids Engineering Division of ASME for publication in the
JOURNAL OF FLUIDS ENGINEERING. Manuscript received June 26, 2008; final manuscript
received January 19, 2009; published online April 1, 2009. Assoc. Editor: Chunill
Hah.
Waterjet Pumps
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
discharge valve
air
water
mixed-flow pump
redistribution device
and pitot tubes
The nonuniform inflow velocity distribution at normal operating condition is a phenomenon, which is widely recognized
2224.
As an example, measured axial velocity distributions are shown
in Fig. 2 25. Two operating conditions are given, representing a
fully loaded and an unloaded vessel at maximum power. Operating conditions of waterjet propelled ships are customarily classified according to the inlet velocity ratio IVR, which is defined as
IVR = vs/v p
with vs the ship speed and v p the average axial velocity entering
the pump. Thus, a fully loaded ship at maximum velocity has a
medium IVR value while an unloaded ship at maximum velocity
is characterized by a high value of IVR. Operating conditions
during the start-up phase are characterized by low values of IVR.
Experimental Setup
Dynamometer
thrust
bearing
radial
bearing
discharge pipe
double roller
bearing
main
shaft
stator
secondary
shaft
dynamometer
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
1.0
uniform
0.8
profile A
profile B
y [-] 0.6
0.4
0.2
0
0.0
0.5
1.0
1.5
2.0
v / vaverage [-]
Since the axial velocity profile in the straight suction pipe of the
setup is approximately uniform, a flow redistribution device of
some sort is required. Its purpose is to modify the velocity distribution into a nonuniform profile as given in Fig. 2. A bundle of
small diameter tubes was shaped according to the method of Kotansky 26, extended to three dimensions. However, measurements showed that the resulting velocity distribution, though
stable and nonuniform, did not resemble the intended shear profile. Velocity measurements were performed with Pitot tubes in a
Results
Experiments are performed with and without velocity redistribution device for different shaft speeds ranging from
450 rpm to 900 rpm. Reynolds numbers are in the range Re
= 5 11 106, and Rev = 6 12 105, based on BEP operation.
Flow rates are varied between 10% QBEP and the maximum flow
rate of 120% QBEP 110% QBEP with pipe bundle installed. For
each run, bridge signals are sampled during 500 shaft revolutions and ensemble averaged. No additional filtering is performed
apart from the 50 Hz line frequency.
In Sec. 6.1 global characteristics of the pump are given for both
uniform and nonuniform suction flows. Results of force measurements are given in Sec. 6.2, for a shaft speed of 700 rpm at BEP
flow rate. The influence of the nonuniform suction flow distribution, as well as blade interaction on radial forces, is shown. A
detailed study of the influence of flow rate on the steady fluidinduced force is presented in Sec. 6.3.
6.1 Global Performance Characteristics. Figures 911
present characteristic curves for manometric head, torque, and
axial force, at different shaft speeds. No suction flow redistribution device was installed. The manometric head was calculated
1.0
medium IVR
0.8
high IVR
profile B
y [-] 0.6
0.4
0.2
0
0.0
0.5
1.0
1.5
2.0
v / vaverage [-]
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
3.0
3.0
900 rpm
2.5
2.5
750 rpm
2.0
gHs / (N2D2)
gHm / (N2D2)
700 rpm
2.0
1.5
uniform
750 rpm
800 rpm
1.0
non-uniform
small clearance
large clearance
1.5
0.5
0.7
0.9
2.00
1.0
1.75
0.5
0
0.5
1.50
-0.5
0
0.0
0.4
0.6
0.8
1.0
900 rpm
800 rpm
750 rpm
700 rpm
T / (N2D5)
0.14
0.12
0.10
0.08
0.4
0.6
0.8
1.0
1.2
Q / QBEP
1.0
1.2
900 rpm
0.16
700 rpm
uniform
750 rpm
750 rpm
1.1
non-uniform
0.15
1.0
0.14
T / (N2D5)
Fax / (N2D4)
0.8
800 rpm
0.9
0.13
0.12
0.8
small clearance
0.11
0.7
0.0
0.6
Q / QBEP
Q / QBEP
0.2
0.4
1.2
0.0
0.2
0.0
0.2
large clearance
0.10
0.2
0.4
0.6
0.8
1.0
1.2
Q / QBEP
0.0
0.2
0.4
0.6
0.8
1.0
1.2
Q / QBEP
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
0.5
small clearance
750 rpm
0.5
uniform
non-uniform
0.0
700 rpm
100% Q
BEP
FY / Fg
0.4
large clearance
static efficiency
0.3
-0.5
-1.0
-1.5
0.2
-2.0
-1.5
-1.0
-0.5
0.1
0.0
0.5
1.0
1.5
FX / Fg
-0.1
0.2
0.4
0.6
0.8
1.0
1.2
Q / QBEP
tates in the direction of shaft revolution. Time-averaged hydrodynamic force is small. Rotor-stator blade interaction force is shown
in Fig. 18, in the inertial frame of reference. Note that this blade
passing excitation force induces a backward whirling motion of
the impeller at a frequency of six times shaft frequency.
In case of a nonuniform suction flow distribution, similar force
plots are obtained. However, an additional steady fluid-induced
force is observed in a direction, which depends on the suction
flow profile. A detailed analysis is given in Sec. 6.3.
6.3 Steady Fluid-Induced Forces. Hydrodynamic forces
were measured for different combinations of shaft speed and flow
rate, and for different suction flow profiles. In case of nonuniform
suction flow, a steady fluid-induced force emerges, which results
from an unbalanced torque loading of the impeller blades 27.
The magnitude and direction of this steady force are given in Fig.
19, for different flow rates. It is apparent that the force magnitude
increases with flow rate, which confirms the findings of Bolleter et
0.5
0.0
FY / Fg
0.0
700 rpm
100% Q
BEP
-0.5
-1.0
-1.5
1.3
-2.0
uniform
750 rpm
-1.5
non-uniform
-1.0
-0.5
0.0
0.5
1.0
1.5
FX / Fg
1.2
1.1
small clearance
1.0
1.0
0.5
FY / Fg
Fax / (N2D4)
0.9
large clearance
0.8
0.0
-0.5
-1.0
-1.5
0.7
0.0
700 rpm
100% Q
BEP
0.2
0.4
0.6
0.8
1.0
1.2
Q / QBEP
-1.5
-1.0
-0.5
0.0
0.5
1.0
1.5
FX / Fg
Fig. 18 Force vector plot for frequencies above shaft frequency showing rotor-stator blade interaction and backward
whirl of the impeller
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
106%QBEP
0.02
900 rpm
800 rpm
profile A
700 rpm
profile B
625 rpm
0.03
600 rpm
96%
0.01
FY / (N2D4)
0.04
uniform
700 rpm
118%
0.00
F / (N2D4)
0.03
74%
68%
50%
106%
450 rpm
0.00
-0.02
-0.02
500 rpm
0.02
0.01
82%
-0.01
550 rpm
-0.01
0.00
0.01
0.02
0.03
0.0
0.2
FX / (N2D4)
0.4
0.6
0.8
1.0
1.2
Q / QBEP
al. 20. The direction of the force is opposite for the two nonuniform suction flow profiles A and B, consistent with the two profiles being approximately inversed in vertical direction.
Further measurements for suction profile B indicative of the
suction flow profile in practical waterjet applications are presented in Figs. 2022, for different shaft speeds. Note that the
steady fluid-induced force scales with the square of the shaft
speed. The direction of the force vector with respect to the horizontal axis is 35 45 deg, for flow rates above 50% QBEP. For
lower flow rate the direction deviates, presumably due to inlet
recirculation, which influences the suction velocity profile.
lection of the drive train components should then allow for adjustment of the intended shaft speed in order to compensate for the
reduced head.
A second phenomenon related to nonuniform suction flow is the
radial loading of the impeller. It is of interest to compare the
magnitude of the forces measured in this investigation to forces
mentioned in literature for pumps of the same type. An overview
of force measurements on pump shafts can be found in, e.g., Ref.
28. The magnitude of forces is customarily divided into steady
and unsteady components in the inertial frame of reference. Figure 23 presents a comparison of the steady force components, as a
function of flow rate. Clearly, at a larger flow rate, the measured
force is much higher than the values normally anticipated in these
pumps. A high radial loading not accounted for will lead to excessive wear of bearings and seals and even rubbing of the impeller
at the casing wall. Moreover, the force being steady in the inertial
frame of reference is potentially harmful to the shaft, since it may
lead to damage through fatigue failure.
A second remark on the radial loading of the impeller concerns
the direction of the force vector. From the measurements it was
concluded that the impeller undergoes a radial force with an upward component. Depending on the type of suction velocity profile this force may be in upward direction entirely. For example,
consider a double-suction pump with its suction nozzle in an unfavorable direction, or a waterjet system in a ship during maneu-
An important observation in this study is that the global performance of a pump is influenced by the type of inflow velocity
profile. For the relatively mild nonuniformity in the experiments,
the head, torque, and axial force were reduced by a few percent.
And although the efficiency of the pump was not reduced, it still
has an important practical implication. Many pumping systems
that face nonuniform inflow conditions during service are selected
based on performance under uniform inflow, e.g., sump pumps or
cooling water pumps. The reason is that most pump manufacturers
use standard closed-loop test-rigs with straight suction piping. Se0.03
900 rpm
140
800 rpm
700 rpm
450 rpm
0.00
625 rpm
600 rpm
550 rpm
0.01
700 rpm
100
angle [degr]
FY / (N2D4)
600 rpm
500 rpm
800 rpm
120
625 rpm
0.02
900 rpm
80
550 rpm
500 rpm
60
450 rpm
40
20
0
-0.01
-0.01
0.00
0.01
0.02
0.03
FX / (N2D4)
-20
0.0
0.2
0.4
0.6
0.8
1.0
1.2
Q / QBEP
Fig. 22 Direction of steady fluid-induced force vector for different shaft speeds and suction velocity profile B
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
0.05
900 rpm
800 rpm
700 rpm
0.04
625 rpm
600 rpm
F / (N2D4)
550 rpm
500 rpm
0.03
450 rpm
0.02
Glich (mixed-flow,
vaned casing)
0.01
0.00
0.0
0.2
0.4
0.6
0.8
1.0
1.2
Q / QBEP
Acknowledgment
The author would like to thank Wrtsil Propulsion Netherlands
for providing the scale-model pump and supporting the publication of this paper.
Nomenclature
gravitational acceleration m / s2
impeller outer diameter m
force N
manometric pump head m
static pump head m
shaft speed rps
flow rate m3 / s
torque N m
Reynolds number based on impeller tip speed,
Re = D2 / v
Rev Reynolds number based on pipe velocity and
diameter
kinematic viscosity m2 / s
mass density kg/ m3
g
D
F
Hm
Hs
N
Q
T
Re
References
1 Ehrich, F. F., and Childs, D. W., 1984, Self-Excited Vibrations in High Performance Turbomachinery, Mech. Eng. Am. Soc. Mech. Eng., 106, pp.
6679.
2 Agostinelli, A., Nobles, D., and Mockridge, C. R., 1960, An Experimental
Investigation of Radial Thrust in Centrifugal Pumps, ASME J. Eng. Power,
82, pp. 120126.
3 Iversen, H. W., Rolling, R. E., and Carlson, J. J., 1960, Volute Pressure
Distribution, Radial Force on the Impeller and Volute Mixing Losses of a
Radial Flow Centrifugal Pump, ASME J. Eng. Power, 82, pp. 136144.
4 Domm, U., and Hergt, P., 1970, Radial Forces on Impeller of Volute Casing
Pumps, Flow Research on Blading, L. S. Dzung, ed., Elsevier, Netherlands,
pp. 305321.
5 Hergt, P., and Krieger, P., 1969, Radial Forces in Centrifugal Pumps With
Guide Vanes, Proc. Inst. Mech. Eng., 184, pp. 101107.
6 Uchida, N., Imaichi, K., and Shirai, T., 1971, Radial Force on the Impeller of
a Centrifugal Pump, Bull. JSME, 14, pp. 11061117.
7 Chamieh, D., Acosta, D. S., Brennen, C. E., and Caughey, T. K., 1985, Experimental Measurements of Hydrodynamic Radial Forces and Stiffness Matrices for Centrifugal Pump Impeller, ASME J. Fluids Eng., 1073, pp. 307
315.
8 Jerry, B., Acosta, A. J., Brennen, C. E., and Caughey, T. K., 1985, Forces on
Centrifugal Pump Impellers, Proceedings of the second International Pump
Symposium, Houston, TX, Apr. 30May 2.
9 Guinzburg, A., Brennen, C. E., Acosta, A. J., and Caughey, T., 1993, The
Effect of Inlet Swirl on the Rotordynamic Shroud Forces in a Centrifugal
Pump, ASME J. Eng. Gas Turbines Power, 115, pp. 287293.
10 Uy, R. V., and Brennen, C. E., 1999, Experimental Measurements of Rotordynamic Forces Caused by Front Shroud Pump Leakage, ASME J. Fluids
Eng., 121, pp. 633637.
11 Hsu, Y., and Brennen, C. E., 2002, Effect of Swirl on Rotordynamic Forces
Caused by Front Shroud Pump Leakage, ASME J. Fluids Eng., 124, pp.
10051010.
12 Ohashi, H., Imai, H., and Tsuchihashi, T., 1991, Fluid Force and Moment on
Centrifugal Impellers in Precessing Motion, ASME FED Fluid Machinery
Forum, Vol. 1193, Portland, OR, pp. 5760.
13 Yoshida, Y., Tsujimoto, Y., Ohashi, H., and Kano, F., 1999, The Rotordynamic Forces on an Open-Type Centrifugal Compressor Impeller in Whirling
Motion, ASME J. Fluids Eng., 121, pp. 259265.
14 Yoshida, Y., Tsujimoto, Y., Yokoyama, D., Ohashi, H., and Kano, F., 2001,
Rotordynamic Fluid Force Moments on an Open-Type Centrifugal Compressor Impeller in Precessing Motion, Int. J. Rotating Mach., 7, pp. 237251.
15 Yoshida, Y., Tsujimoto, Y., Morimoto, G., Nishida, H., and Morii, S., 2003,
Effects of Seal Geometry on Dynamic Impeller Fluid Forces and Moments,
ASME J. Fluids Eng., 125, pp. 786795.
16 Suzuki, T., Prunires, R., Horiguchi, H., Tsukiya, T., Taenaka, Y., and
Tsujimoto, Y., 2007, Measurements of Rotordynamic Forces on an Artificial
Heart Pump Impeller, ASME J. Fluids Eng., 129, pp. 14221427.
17 Toyokura, T., 1961, Studies on the Characteristics of Axial-Flow Pumps,
Bull. JSME, 4, pp. 287340.
18 Badowski, H. R., 1970, Inducers for Centrifugal Pumps, Worthington
Canada, Ltd., Internal Report.
19 Del Valle, D., Braisted, D. M., and Brennen, C. E., 1992, The Effects of Inlet
Flow Modification on Cavitating Inducer Performance, ASME J. Fluids Eng.,
114, pp. 360365.
20 Bolleter, U., Leibundgut, E., Sturchler, R., and McCloskey, T., 1989, Hydraulic Interaction and Excitation Forces of High Head Pump Impellers, Joint
ASCE/ASME Pumping Machinery Symposium, San Diego, CA, pp. 187193.
21 Fujii, A., Azuma, S., Yoshida, Y., Tsujimoto, Y., and Laffite, S., 2002, Unsteady Stress of 4-Bladed Inducer Blades and the Effect of Inlet Flow Distortion, JSME Int. J., Ser. B, 45, pp. 4754.
22 Seil, G. J., Fletcher, C. A., and Doctors, L. J., 1995, The Application of
Computational Fluid Dynamics to Practical Waterjet Propulsion System Design and Analysis, Proceedings FAST95, pp. 13791389.
23 Verbeek, R., and Bulten, N. W. H., 1998, Recent Developments in Waterjet
Design, Proceedings Rina Waterjet Propulsion Conference, Amsterdam.
24 Brandner, P., and Walker, G. J., 2001, A Waterjet Test Loop for the Tom Fink
Cavitation Tunnel, Proceedings Waterjet Propulsion III Conference, Gothernburg.
25 Bulten, N. W. H., Verbeek, R., and van Esch, B. P. M., 2006, CFD Simulations of the Flow Through a Waterjet Installation, Int. J. Marit. Eng., 148, pp.
2333.
26 Kotansky, D. R., 1966, The Use of Honeycomb for Shear Flow Generation,
AIAA J., 4, pp. 14901491.
27 Bulten, N. W. H., and van Esch, B. P. M., 2007, Fully Transient CFD Analyses of Waterjet Pumps, Marine Technology, Vol. 443, pp. 185193, Society
of Naval Architects and Marine Engineers SNAME.
28 Glich, J. F., 1999, KreiselpumpenEin Handbuch fr Entwicklung, Anlageplanung und Betrieb, Springer-Verlag, Berlin.
29 Dubas, M., 1984, ber die Erregung infolge der Periodizitt von Turbomachinen, Ing.-Arch., 54, pp. 413426.
30 Brennen, C. E., 1994, Hydrodynamics of Pumps, Oxford University Press,
Oxford.
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
David L. Bell
Alstom Power Ltd.,
Rugby CV21 2NH, UK
This paper presents a combined experimental and computational study of unsteady flows
in a linear turbine cascade oscillating in a three-dimensional bending/flapping mode.
Detailed experimental data are obtained on a seven-bladed turbine cascade rig. The
middle blade is driven to oscillate and oscillating cascade data are obtained using an
influence coefficient method. The numerical simulations are performed by using a 3D
nonlinear time-marching NavierStokes flow solver. Single-passage domain computations
for arbitrary interblade phase angles are achieved by using the Fourier shape correction
method. Both measurements and predictions demonstrate a fully 3D behavior of the
unsteady flows. The influence of the aerodynamic blockage introduced by part-span
shrouds on turbine flutter has been investigated by introducing flat plate shaped shrouds
at 75% span. In contrast to practical applications, in the present test configuration, the
mode of vibration of the blades remains unchanged by the introduction of the part-span
shroud. This allows the influence of the aerodynamic blockage introduced by the partspan shroud to be assessed in isolation from the change in mode shape. A simple shroud
model has been developed in the computational solver. The computed unsteady pressures
around the shrouds are in good agreement with the experimental data, demonstrating the
validity of the simple shroud model. Despite of notable variations in local unsteady
pressures around the shrouds, the present results show that the blade aerodynamic damping is largely unaffected by the aerodynamic blockage introduced by part-span shrouds.
DOI: 10.1115/1.3111254
Introduction
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Experimental Study
X
mm
Y
mm
0.00
0.11
0.48
0.74
2.25
6.16
9.91
13.53
16.16
19.58
22.89
26.14
29.32
32.44
35.49
38.49
41.43
44.32
47.15
49.93
52.65
55.32
57.94
60.50
63.01
65.46
67.85
70.17
72.43
74.61
76.73
78.82
80.85
82.85
84.81
86.73
88.62
90.48
92.29
94.95
97.50
99.56
101.55
103.45
105.28
106.69
108.03
109.32
110.28
110.30
110.29
110.18
109.96
109.66
109.32
108.98
108.68
108.45
108.37
107.99
106.46
104.48
102.40
100.25
98.04
95.80
93.54
91.23
88.86
86.43
83.94
92.26
92.91
93.46
93.67
94.46
96.28
97.69
98.76
99.37
99.93
100.26
100.41
100.38
100.20
99.87
99.38
98.76
98.00
97.11
96.08
94.91
93.62
92.19
90.62
88.91
87.05
85.04
82.87
80.53
78.00
75.34
72.56
69.67
66.67
63.59
60.42
57.16
53.82
50.37
45.02
39.44
34.60
29.58
24.37
18.96
14.48
9.85
5.08
1.20
1.02
0.85
0.52
0.25
0.07
0.00
0.05
0.22
0.48
0.64
1.74
5.93
10.94
15.73
20.32
24.79
29.18
33.49
37.70
41.76
45.67
49.42
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Table 1
Continued.
X
mm
Y
mm
81.41
78.88
76.34
73.70
70.96
68.12
65.20
62.19
59.10
55.94
52.70
49.39
46.69
43.25
39.75
36.21
33.33
30.39
27.41
24.37
20.49
16.55
12.60
8.66
4.71
1.49
0.85
0.34
0.05
53.07
56.72
60.34
63.73
66.85
69.73
72.39
74.85
77.10
79.16
81.04
82.73
83.96
85.33
86.56
87.68
88.44
89.08
89.58
89.96
90.24
90.35
90.45
90.55
90.65
90.73
90.91
91.33
91.93
and the hub displacement, profiled slots were located on the hub
wall. It should be mentioned that sealing measure was taken to
prevent flow leakage though these slots.
2.2 Instrumentation, Data Acquisition and Reduction. The
oscillating blade and one nonoscillating blade are instrumented
with pressure tappings at six spanwise sections 10%, 30%, 50%,
70%, 80%, and 95% span sections. At each section, 14 pressure
tappings are located on the suction surface SS and 10 on the
pressure surface PS, as indicated in Fig. 2. In total, there are 144
tappings on each blade surface. These pressure tappings are used
for both steady and unsteady pressure measurements. The instrumented nonoscillating blade is interchangeable with all other stationary blades, thereby enabling both steady and unsteady pressure measurements at all blade positions in the cascade.
In this low-speed test facility, realistic reduced frequencies can
be achieved at low frequencies of blade vibration. For example, a
nominal frequency of 10.3 Hz here is equivalent to a reduced
frequency of 0.4, which is based on the blade chord and exit
isentropic velocity. This makes it possible for the unsteady pressure signals to be recorded with off-board pressure transducers.
The rig is equipped with five off-board pressure transducers Sen-
sym 142C01D, 01 psi range, so a total of 30 sets of measurements are required for each blade in one test. Unsteady signals
from the transducers are recorded on a personal computer PC
through an Amplicon PC30G data-logging card. The unsteady
pressure acquisition procedure is synchronized by a signal from
an optical Schmitt trigger.
The unsteady pressure signals are ensemble-averaged over 150
periods. The ensemble-averaged unsteady pressure signals are reduced into their harmonic components using a Fourier series. To
correct the unsteady pressure signal for phase shift and attenuation, along the tubing lengths, a tubing transfer function TTF
scheme, proposed by Irwin et al. 13, is applied. The present
application of the TTF method follows the procedures of SimsWilliams 14 and Yang and He 9.
Once the unsteady pressure measurements are obtained, on
each individual blade, they are superposed to construct the tuned
cascade data by utilizing the influence coefficient method. The
validity of the method for this case has been validated by examining the convergence of influence coefficients on those blades
away from the central oscillating blade and the linearity of the
unsteady pressures at different oscillation amplitudes. As described earlier, all influence coefficients have been reduced into
their harmonic components. Accordingly, the first harmonic component Cp1 for a tuned cascade can be expressed as a sum of the
first harmonic components Cp1n of influence coefficients from
all seven blades:
3
Cp1 =
Blade chord, C
Blade span, h
Pitch length, S
Stagger angle
Inlet flow angle
Vibration mode
Bending amplitude
Reynolds number, Re
Reduced frequency, k
Nominal frequency, f
0.143 m
0.180 m
0.100 m
40.0 deg
0.0 deg
First bending normal to absolute chord
3%C at tip, 0.3%C at hub
2.2 105
0.2, 0.4, 0.6
5.2 Hz, 10.3 Hz, 15.5 Hz
in
n=3
where n is the blade index number, and is the inter blade phase
angle. The direction of blade indexing is the same as that of a
forward traveling wave i.e., that of the rotation if the cascade was
taken from a turbine rotor with index 0 denoting the middle reference blade. Consequently, the sign convention for the interblade
phase angle IBPA is such that a negative IBPA corresponds to a
backward traveling wave mode.
2.3
Cp ne
Experimental Results
2.3.1 Steady Flow. The steady flow results are discussed aiming at identifying the main steady flow features. Since the major
unsteady pressure response happens on the central three blades,
which will be verified later, the steady flow results are, therefore,
presented only on the central three blades.
Shown in Figs. 3a3c are the steady flow blade surface pressure distributions for the three middle blades at the flow conditions described in Table 2. The steady flow blade surface pressure
measurements were obtained at six spanwise sections from 10%
to 95% span. For each blade, the blade surface pressure distributions are almost identical for different spanwise sections, although
slight deviations in pressure distribution are observed toward
MAY 2009, Vol. 131 / 051102-3
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
blade ends over the suction surface. These deviations are believed
to be associated with the endwall effects with slight unloading
present on the suction surface. These slight endwall effects are
attributable to the favorable pressure gradient present in the cascade and the relatively small blade deflection angle. Overall, the
results indicate that the steady flow blade surface pressure response is predominantly two dimensional.
Figure 3d shows the comparison of the steady flow blade
surface pressure distributions at the midspan section on the three
central blades. Because of the predominant two-dimensional feature of the steady flow, this comparison at the midspan section is
sufficient to verify the blade-blade periodicity in the cascade. The
plot demonstrates an excellent comparison on the pressure surface, while a good comparison is achieved on the suction surface
despite some deviation present on the rear half chord of blade 1.
Good blade-blade periodicity achieved in the steady flow surely
forms a sound basis for the unsteady flows presented hereafter.
2.3.2 Unsteady Flow. To assess the uncertainty of unsteady
flow measurements, a series of tests was performed to evaluate the
experimental errors and repeatability. In these tests, five tappings
were chosen from the 70% span of the oscillating blade to provide
a representative range of unsteady pressure response. For each
tapping, 40 sets of measurements were obtained, through the ex-
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Fig. 6 Experimental test for linearity: first harmonic pressure coefficient at 50% span
section on blade 0 at two bending amplitudes k = 0.4
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Fig. 8 Amplitudes of first harmonic pressure at IBPA= 60 deg: a suction surface and b
pressure surface
Computational Study
Fig. 9 Phase angles of first harmonic pressure at IBPA= 60 deg: a suction surface b and pressure surface
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Fig. 10 Global aerodynamic damping at three reduced frequencies bold line: trend of the least stable IBPA with regard
to reduced frequency
NFou
NFou
x, ,r,t =
U
A x, ,rsinnt + B x, ,rcosnt
n
n=1
where An and Bn are the nth order Fourier coefficients; NFou is the
order of the Fourier series; x, , and r are the axial, circumferential, and radial coordinates, respectively, and t is the physical time.
The phase-shifted periodic conditions are then defined as
x, G,r,t =
U
A x, ,rsinnt
n
n=1
+ Bnx, ,rcosnt
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Figs. 13 and 14. Discrepancy between the predicted and measured amplitudes of the first harmonic pressure is observed at
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
The shroud modeling is achieved by applying solid wall boundary conditions at shroud surface locations. Figure 18 shows the
computational mesh used for the study of effects of part-span
shrouds. It consists of 132 nodes in the axial direction, 41 in the
pitchwise direction and 51 in the spanwise direction. The mesh
clustering is applied toward the shroud location in the spanwise
direction and finer mesh spacing is also used around the shroud
leading and trailing edges. For the comparison purpose, the same
mesh density is utilized in both calculations with and without
shrouds to eliminate the influence of mesh dependency on the
results. During blade oscillation, the shrouds are assumed to
stretch in a plane of constant span height. In other words, the
analyzed shroud configuration is otherwise identical to the tested
one except that the radial movement of shrouds is neglected. The
part-span shroud plates are thin and do not induce noticeable
steady pressure loading change on the blade surface. The results
presented are, therefore, confined to those for the unsteady flow.
Figure 19 shows the amplitudes of the first harmonic pressure at
50%, 70%, 80%, and 95% span sections for settings with and
without part-span shrouds. For each section, the left plot shows
the experimental results and the right shows the predictions. Upon
inspection of the set of the first harmonic pressure, it is found that
the influence of part-span shrouds manifests itself mainly in terms
of the change in the pressure amplitude on the suction surface.
The presence of the part-span shrouds results in reduced ampli-
tudes for the sections below 75% span and increased amplitudes
for those sections above 75% span. It is evident that the computational predictions for settings with and without part-span
shrouds reveal exactly the same behavior as that observed in the
measurement. This demonstrates the overall validity of the present
simple shroud model.
The opposite trends in the amplitude change of the first harmonic pressure for spanwise sections below and above the partspan shrouds reflect a 3D behavior of the unsteady pressure propagation. Recall the observation on the results in Fig. 8 that the
amplitude of the first harmonic pressure is nonproportional to the
local blade vibration amplitude along the blade span. The basic
mechanism is that the spanwise unsteady interaction results in an
instantaneous redistribution of the unsteady pressure. The unsteady loading is alleviated toward the tip and enhanced toward
the hub, which can be simply regarded as an unsteady 3D relief.
The presence of part-span shrouds, on the other hand, blocks a
radial propagation of unsteady pressures and hence reduces the
3D unsteady relief effects. Consequently, the amplitude of the first
harmonic pressure with part-span shrouds is larger than that of the
nominal case for sections above 75% span, whereas the trend in
the amplitude change is reversed for sections below 75% span.
Figure 20 shows the phase angles of first harmonic pressure for
settings with and without part-span shrouds. Regardless the spanwise locations, the phase angles of the unsteady pressure are
largely unaffected.
The overall influence of the part-span shrouds on the blade
flutter is indicated by the integrated aerodynamic damping. The
experimental data show that the part-span shroud has a small stabilizing effect, and this behavior has been also correctly captured
by the CFD results. The differences in the aerodynamic damping
as caused the stabilizing effects are, however, comparable to or
smaller than the discrepancies between the experimental data and
the CFD predictions for the nominal case as shown in Fig. 16.
Hence, the overall influence of the part-span shroud can be regarded as insignificantly small for the present case studied. This is
largely due to the fact that the part-span shrouds have predominantly local effects on unsteady pressures and hence the aerodynamic damping.
5
Thickness
Length in axial direction
Spanwise location
Axial location
3 mm
60 mm
75% span
18 78%Cax
Concluding Remarks
A linear oscillating turbine cascade is investigated experimentally and computationally. Detailed unsteady pressure measurements are made at six spanwise sections from 10% to 95% blade
height. The steady pressure coefficient distributions reveal a predominant 2D steady background flow, while the unsteady pressure
results demonstrate a strong 3D behavior of unsteady pressure
response to a three-dimensional blade flapping oscillation. The
computational solutions are obtained from a 3D unsteady timemarching NavierStokes flow solver using a single-passage domain. The computational results are in good agreement with experimental data.
A further study is performed to investigate the influence of the
aerodynamic blockage that is introduced by the application of
part-span shrouds, on the blade flutter characteristics in the linear
turbine cascade with the same blade oscillation mode shape. The
part-span shrouds are positioned at 75% span height. Notable
variations in local unsteady pressure at spanwise sections near the
shroud location, as shown in the experimental data, are well captured by the computational method with a simple part-shroud
model. The influence of the aerodynamic blockage associated with
the part-span shrouds on the overall aerodynamic damping, however, seems insignificant for the present cascade.
MAY 2009, Vol. 131 / 051102-9
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Fig. 19 Effects of part-span shrouds on amplitude of the unsteady pressure response. Left: experiment. Right: calculation k = 0.4, = 60 deg: a 50% span section, b 70% span section, c 80% span section, and d 95% span section.
Fig. 20 Phase angle of first harmonic pressure for settings with and without partspan shrouds: a 50% span section, b 70% span section, c 80% span section, and
d 95% span section
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Acknowledgment
This work was carried out at University of Durham, Durham
UK, sponsored partially by ALSTOM Power Ltd, which is gratefully acknowledged. The authors would also like to thank Dr.
Peter Walker ALSTOM for continued interest and technical discussions
Nomenclature
Ax bending amplitude, nondimensionalized with
chord
Axtip bending amplitude at blade tip, nondimensionalized with chord
Ap1 amplitude of the first harmonic pressure, Pa
C blade chord length
Cax axial blade chord length
Cp blade surface pressure coefficient, Cp = P
P2 / P01 P2
Cp1 amplitude of the first harmonic pressure coefficient, Cp1 = Ap1 / P01 P2Axtip
h blade span, m
k reduced frequency, k = C / Vref
P2 exit static pressure
P01 inlet stagnation pressure
Re Reynolds number
S pitch length
Vref reference isentropic exit velocity, m/s, Vref
= 2P01 P2 /
x axial coordinate, m
circumferential coordinate, deg
phase angle, deg
IBPA, deg
angular frequency, rad/s
aerodynamic damping coefficient, = 1 / hh0ldr
l local aerodynamic damping coefficient,
l = C AxlCp1sin1 / C Axtipds
Subscript
References
1 Gerolymos, G. A., 1993, Advances in the Numerical Integration of the ThreeDimensional Euler Equations in Vibrating Cascades, ASME J. Turbomach.,
115, pp. 781790.
2 Hall, K. C., and Clark, W. S., 1993, Linearized Euler Predictions of Unsteady
Aerodynamic Loads in Cascades, AIAA J., 31, pp. 540550.
3 He, L., and Denton, J. D., 1994, Three-Dimensional Time-Marching Inviscid
and Viscous Solutions for Unsteady Flows Around Vibrating Blades, ASME
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Seyed Mohammad
Hosseini
Department of Energy,
Division of Fluid Mechanics,
Lund University,
Lund 22100, Sweden
e-mail: seyed_mohammad.hosseini@energy.lth.se
Kazuhisa Yuki
Hidetoshi Hashizume
Department of Quantum Science and Energy,
Tohoku University,
6-6-01 Aza-Aoba, Aramaki,
Sendai 980-8579, Japan
Experimental Investigation of
Flow Field Structure in Mixing
Tee
T-junction is one of the familiar components in the cooling system of power plants with
enormous capability of high-cycle thermal fatigue. This research investigates the structure and mixing mechanism of turbulent flow in a T-junction area with a 90 deg bend
upstream. According to the wide distribution of turbulent jets in the T-junction, a reattached jet was selected previously as the best representative condition with the highest
velocity fluctuation and the most complex structure. For considering the mixing mechanism of re-attached jet, T-junction is subdivided into few lateral and longitudinal sections, and each section is visualized separately by particle image velocimetry technique.
Corresponding to the experimental data, the branch flow acts as a finite turbulent jet,
develops the alternative type of eddies, and causes the high velocity fluctuation near the
main pipe wall. Three regions are mainly subject to maximum velocity fluctuation: the
region close to the jet boundaries (fluctuation mostly is caused by KelvinHelmholtz
instability), the region above the jet and along the main flow (fluctuation mostly is caused
by Karman vortex), and the re-attached area (fluctuation mostly is caused by changing
the pressure gradient in the wake area above the jet). Finally, the re-attached area (near
the downstream of wake area above the jet) is introduced as a region with strongest
possibility to high-cycle thermal fatigue with most effective velocity fluctuation on the
main pipe wall above the branch nozzle. DOI: 10.1115/1.3112383
Keywords: fluid mixing structure and interaction, mixing tee, PIV, piping system, turbulent flow
Introduction
This research investigates fluid mixing phenomena and turbulent jet structure in a T-junction area with a 90 deg bend upstream
to understand the basic mechanism of high-cycle thermal fatigue.
The problem of thermal fatigue occurs in the pipes where two
flows with different temperatures mix together. The T-junction is
one of the familiar components with a considerable potential of
thermal fatigue and is used in many thermohydraulic systems such
as combustion engines, turbines, exhaust systems, hydraulicspneumatics, and reheat systems. With due attention to the importance of thermal striping phenomenon in the T-junction area, several experiments and analyses have been performed such as the
evaluation of the thermal fatigue by Faidy 1, the numerical
simulation of the mixing phenomenon by Tanaka 2, the analysis
of the flow field structure by Igarashi 3, and the LES study of the
high-cycle temperature fluctuation by Hu and Kazimi 4. These
researches consider the T-junction as a single component. Since
usually the T-junction is connected to other apparatuses in the
piping systems, the 90 deg bend is chosen as one of the common
components in most piping systems and it was also connected in
the upstream of the T-junction in some familiar leakage accidents
of power plants such as Phenix and Civaux in France. A secondary
flow is formed by this 90 deg bend and has strong effects on the
flow field along with the mixing mechanism. These effects were
studied previously by Hosseini and Yuki 57. The interaction
between main and branch flows including the effects of secondary
flow formed an unstable area above the branch nozzle. This area
was shown to be highly loaded and subjected to the temperature
Contributed by the Fluids Engineering Division of ASME for publication in the
JOURNAL OF FLUIDS ENGINEERING. Manuscript received February 13, 2008; final manuscript received February 25, 2009; published online April 13, 2009. Assoc. Editor:
Juergen Kompenhans. Paper presented at the 2007 ASME Fluids Engineering Division Summer Meeting and Exhibition FEDSM2007, San Diego, CA, July 30
August 2, 2007.
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
initial and transient parts of jet causes KelvinHelmholtz instability and creates the high vorticity field in the same area. In addition
there are many other eddies in the downstream of T-junction that
mainly are created by the following mechanisms: the oscillation
of the flexible body of jet in the transitional part caused by the
frequent drag and lift forces, the structure-interaction between the
flow and piping system, the secondary flow created by 90 deg
bend in the upstream, the self-developing behavior of turbulent jet
in the main part, etc. Finally, the T-junction area was subdivided
into few regions based on their intensity of velocity fluctuation
near the wall, and the turbulent behavior of the most intense region was analyzed in more detail along with the fluctuation
mechanism.
Experimental Apparatuses
The particle image velocimetry PIV system is used to visualize the flow field in the T-junction area with water as the working
fluid. In this system, the velocity vectors are derived from subsections of the target area of the particle-seeded flow by measuring
the movement of particles between two light pulses. The flow is
illuminated in the target area with a light laser sheet. The camera
is able to capture each light pulse in separate image frames. Once
a sequence of two light pulses is taken, the images are divided
into small subsections called interrogation areas. The interrogation
areas from each image, frames 1 and 2, are cross correlated with
each other pixel by pixel. The correlation produces a signal peak,
identifying the common particle displacement. An accurate measure of the displacementand thus also the velocityis achieved
with subpixel interpolation. A velocity vector map over the whole
target area is obtained by repeating the cross-correlation for each
interrogation area over the two image frames captured by the
charge coupled device CCD camera. The correlation of the two
interrogation areas, frames 1 and 2, results in the particle displacement represented by a signal peak in the correlation.
The number of particles in the flow is important for obtaining a
good signal peak in the cross-correlation. As a rule of thumb,
1025 particle images should be seen in each interrogation area.
When the size of the interrogation area, the magnification of the
imaging, and the light-sheet thickness are known, the measurement volume can be defined. Setting up a PIV measurement, the
side length of the interrogation area d and the image magnification
sm are balanced against the size of the flow structures to be resolved. One way of expressing this is to require the velocity gradient to be small within the interrogation area:
sm . Umax Umin t
5%
d
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
1
n
2
r,j
2
+ uz,i
1
sj =
u j,i u j2
n i=1
In this figure the different parts of the flow field are numbered,
and the experiment related to each part is described in the following paragraphs. To have comprehensive evaluation of every part,
the T-junction area is subdivided in the several lateral and longitudinal sections A, B, C1, C2, C3, and C4, as shown in Fig. 3,
and each section is visualized separately with high accuracy. All
these experiments are done under the re-attached jet condition
1/2
2
Umix = U2b + Um
I = sr + sz/2/Umix
Here, ur,i and uz,i represent the instantaneous radial and axial
velocities at the i frame, Uave is the averaged absolute velocity,
and n is the total number of the frames. Ub and Um represent the
branch and main velocities, respectively, and sr and sz are the
standard deviations of the velocity variation in the radial and axial
directions. I represents the intensity of velocity fluctuation.
The mixing mechanism of two fluids in the T-junction is researched by visualizing the flow field instantaneously, and the
schematic structure of this mixing is drawn in Fig. 2.
Journal of Fluids Engineering
Fig. 3 Positions of lateral and longitudinal sections to visualize the flow field
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
with almost same branch and main velocities and different Reynolds numbers Re= 53,000 and 10,000, respectively. The complexity of this condition is the basic purpose for choosing it,
which was studied previously by Hosseini and Yuki 10.
Figure 4 shows the average velocity distribution and the velocity fluctuation in the longitudinal section A with 0.48 m/s main
velocity and 0.47 m/s branch velocity. The different parts of
branch jet are easily recognizable in this figure with almost the
same structure as the one introduced by Abramovich 12. The
first part is the wedge shape region including the potential core
that is called the initial part with the minimum velocity fluctuation
and the highest instability compared with other parts. The second
part of the branch jet is the transitional part that is distributed from
the edge of the initial part to the upright direction along the branch
jet axis. The cross section of this area becomes larger in the downstream, and the jet in this part has a flexible structure against the
main flow. The main part of the jet is the last part and spreads
from the transitional part of the jet to the long downstream. This
part is the most flexible, unstable, and self-developed. There is an
interface that is raised by the steep velocity gradient between the
main flow and the jet in the initial and transient parts. The nature
of the interface is caused by the viscous shearing force of two
fluids with different velocities, and its thickness depends on the
boundary condition in each part. The wake area above the jet has
the most steep velocity gradient and has strong temperature fluctuation effects in the case of nonisothermal mixing fluids, and if
the interface touches the main pipe wall frequently then it can
cause the high-cycle thermal fatigue. The effect of temperature
fluctuation on the wall is studied previously by several research
groups such as Hosseini and Yuki 10, Yuki et al. 11, Metzner
and Wilke 8, and Chapuliot and Gourdin 9.
Figure 5 presents the average velocity distribution in four longitudinal sections C1, C2, C3, and C4. The main flow is distributed almost symmetrically in all sections, and the cross section of
the branch jet and the wake area above it become more visible by
getting closer radially to the branch pipe nozzle from section C4
to section C1. The intensity of velocity fluctuation is calculated in
the same longitudinal sections, as drawn in Fig. 6. The fluctuation
is minor in the region with highest velocity because of instabilitys
051103-4 / Vol. 131, MAY 2009
law and it is large in the region around the jet along with the edge
of the wake area because of the variation in steep velocity gradient.
Figure 7 shows the velocity distribution and the velocity fluctuation of secondary flow in the lateral section B. The symmetrical
twin vortexindicated by the average velocity vectors in Fig.
7a and basically created by the 90 deg bend in the upstream
has strong effects on the flow field in the mixing area. Three main
high velocity regions are observed as well as a high velocity fluctuation region between twin vortices in the lateral section.
The mixing progress will be monitored better by analyzing every frame of the time series data repeatedly. Figure 8 describes
four continual velocity vector frames with 0.03 s time separation
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Fig. 9 Five continual frames of time series data with close-up visualization
Ub = 0.55 m / s and Um = 0.73 m / s
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
the main flow and the branch jet, and in the case of nonisothermal
mixing condition can cause high-cycle thermal fatigue in the same
area and sometimes inflict considerable damage on the piping
structures. The second region is distributed between the branch
nozzle and the re-attached area in the edge of the wake area, and
the Karman vortices are the main source of this fluctuation. The
frequency of velocity fluctuation in the second region is lower
compared to the first one as well as the velocity gradient. The
third region is the re-attached region with lowest frequency of the
velocity fluctuation. This fluctuation mainly is generated by the
lift and drag force between the main flow and the jet body. Due to
the fact that the jet body is flexible against the main flow, the high
pressure gradient around the jet changes the cross section of the
jet in the transitional part from the circular shape to the oval shape
periodically. Meanwhile the pressure gradient decreases based on
the oval shape and then automatically the cross section of the jet
changes back to the circular shape again, which means that the lift
and drag forces fluctuate side by side. So the jet body starts to
move up and down as well as right and left frequently. This phenomenon causes the low frequency fluctuation of the velocity in
the re-attached area. Because of an efficient thermal transfer to the
structure, this low frequency fluctuation inflicts the most effective
thermal fatigue damage compared with the other two areas and
can be improved simply just by changing the momentum ratio
between the main and the branch flows or the piping geometry,
which is described previously by Hosseini and Yuki 10.
Conclusion
The fluid mixing mechanism is provided experimentally by particle image velocimetry method to investigate the structure of the
turbulent flow field in a T-junction area with a 90 deg bend upstream. For this purpose, several longitudinal and lateral sections
in the mixing area are visualized by the high frame rate of velocity
distribution, and the statistical analyses of time series data give us
the general view of mixing phenomena.
Figure 10 summarizes three regions with the highest velocity
fluctuation in the T-junction. The first one surrounds the branch
nozzle and is created by the KelvinHelmholtz instability between
051103-6 / Vol. 131, MAY 2009
Nomenclature
t
d
Db
Dm
I
Imax
sm
sj
sr
sz
t
U
Uave
Ub
Um
Umax
Umin
ui,j
uj
ur,i
uz,i
z
time separation
side length of the interrogation area
branch pipe diameter
main pipe diameter
intensity of the velocity fluctuation
maximum velocity fluctuation
image magnification
standard deviation in the j direction
radial velocity standard deviation
axial velocity standard deviation
time separation
mean velocity
average of the absolute velocity
branch velocity
main velocity
maximum velocity
minimum velocity
velocity at the i frame in the j direction
average velocity in the j direction
radial velocity at the i frame
axial velocity at the i frame
longitudinal distance along the axial direction
References
1 Faidy, C., 2004, Thermal Fatigue in Mixing Area, Third International Conference on Fatigue of Reactor Components, EPRI-US NRC-OECD NEA,
Seville, Spain.
2 Tanaka, M., 2004, Sixth International Conference on Nuclear Thermal Hydraulic, Operation and Safety, NUTHOS-6, Nara, Japan, ID No. N6P334.
3 Igarashi, M., 2003, Study on Fluid Mixing Phenomena for Evaluation of
Thermal Striping in a Mixing Tee, Tenth International Topical Meeting on
Nuclear Reactor Thermal Hydraulic, NURETH-10, Seoul, Korea.
4 Hu, L. W., and Kazimi, M. S., 2006, LES Benchmark Study of High Cycle
Temperature Fluctuation Caused by Thermal Striping in a Mixing Tee, Int. J.
Heat Fluid Flow, 271, pp. 5464.
5 Hosseini, S. M., and Yuki, H., 2006, Experimental Investigation of Thermal
Hydraulic Characteristics at a Mixing Tee, International Heat Transfer Conference, Sydney, Australia, FCV-17.
6 Hosseini, S. M., and Yuki, H., 2005, Three-Dimensional Study of Flow Mixing Phenomenon, International Conference Nuclear Energy for New Europe,
Bled, Slovenia, September ID: 037.
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
7 Yuki, K., and Hosseini, S. M., 2004, 15th International Conference on Nuclear
Thermal Hydraulic, Operation and Safety NUTHOS-6, Nara, Japan, ID No.
N6P082.
8 Metzner, K. J., and Wilke, U., 2005, European THERFAT Project-Thermal
Fatigue Evaluation of Piping System Tee Connections, Nucl. Eng. Des.,
2352-3, pp. 473484.
9 Chapuliot, S., and Gourdin, C., 2005, Hydro-Thermal-Mechanical Analysis of
Thermal Fatigue in a Mixing Tee, Nucl. Eng. Des., 2355, pp. 5755906.
10 Hosseini, S. M., Yuki, H., and Hashizume, H., 2008, Classification of Turbulent Jets in a T-Junction Area With a 90-deg Bend Upstream, Int. J. Heat Mass
Transfer, 519-10, pp. 24442454.
11 Yuki, K., Sugawara, Y., and Hosseini, S. M., 2008, Influence of Secondary
Flow Generated in a 90-Degree Bend on the Thermal-Hydraulic Characteristics in a Mixing Tee, Nucl. Sci. Eng., 158, pp. 194202.
12 Abramovich, G. N., 1963, The Theory of Turbulent Jets, MIT Press, Cambridge, MA, Chap. 1, p. 3.
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
A. Ahmadi
Department of Mechanical Engineering,
Iran University of Science and Technology-Arak
Branch,
Arak,
Markazi 38181-41167, Iran
e-mail: a_ahmadi@iust.ac.ir
Introduction
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
1 4 4 d
2p
0.8
+ 0.0188 + 0.0063
19,000
Re
0.8
+ 0.04289 1 0.11
+ 0.031
0.94
0.94
0.8
1
1
+ 0.0110.75 2.8
3.5
106
Re
106
Re
0.7
0.3
4
1 4
1.1
D
25.4
1.3
and hence variations in the discharge coefficient have been analyzed. According to the ISO Standards the flow rate is determined
from the differential pressure measured through the orifice plate
1. For a fully developed velocity profile, for a known discharge
coefficient, the mass flow rate can be calculated or for a known
mass flow rate the standard discharge coefficient can be calculated. Nevertheless, the mass flow rate for disturbed flow can be
measured by experiment and then new discharge coefficient Cd
can be calculated from Eq. 2. Thus the percentage shift of the
standard discharge coefficient due to disturbances Cd % can be
determined as
Cd =
Cd Cd0
100
Cd0
P
1 2
u
2
Experimental Facility
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
cone swirler
joint two-piece of the swirler made of two pieces joined
together
two-piece of the swirler with a gap between these two
parts
four-piece of the swirler with a gap between each part
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
bances upstream of the orifice plate. These disturbances can provide a combination of an asymmetric velocity profile and a swirling flow. In order to assess the effect of the swirler flow
conditioner on the disturbed flow, both types of disturbances were
used in the experimental facilities. To achieve an asymmetric velocity profile, block disturbances were used. These blocks were
cut from a circular piece of metal and were placed on the bottom
of the pipe. One block, referred to as the 1/4 disturbance, had a
cross section of 1/4 of the area of the pipe and caused a significant
3.1 Performance of the Different Swirler Flow Conditioner on Disturbed Flow. The results for the two-piece and joint
swirler conditioners are shown in Figs. 6 and 7, respectively. By
comparing the two graphs it can be seen that the change in discharge coefficient caused by each swirler conditioner on its own is
as much as when using a block disturbance. This means that the
swirler flow conditioner can attenuate the effect of an asymmetric
velocity profile. Nevertheless, the effect of these two swirler conditioners on the swirling flow caused by the swirler disturbance is
totally different. As the same time employment of swirler disturbance causes about 3% change in standard C, using joint and
two-piece swirler conditioner for this condition causes about 7%
and more than 10% this line has not been shown in the graph
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
because it is out of scale. Thus, the swirler flow conditioner generally can dampen the effect of block disturbance; however, the
effective swirler flow conditioner on the swirling flow is a main
task. Therefore, other shapes of the swirler have been designed
and their effects on the swirling flow have been tested.
The effect of the four-piece and cone swirler flow conditioner
on the disturbed flow was examined, and the results on the discharge coefficient are presented in Figs. 8 and 9. It can be seen
from Fig. 8 that the four-piece flow conditioner does not have an
attenuating effect on the asymmetry flow caused by the block
disturbance and neither has any better effect on the swirling flow
relative to previous ones. On the other hand, the result of the cone
swirler shown in Fig. 9 illustrates perhaps the best result for
implementation of the swirler flow conditioner on the disturbed
flow. It can be seen that using the cone swirler by its own changes
the discharge coefficient from 4% for low Reynolds numbers to
about 6% for high Reynolds numbers. The change in discharge
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
double elbow disturbances were set in the water rig, and the effect
of these configurations on the standard orifice plate alongside the
cone swirler conditioner was investigated. This result of change in
discharge coefficient is given in Fig. 10.
It can be seen from Fig. 10 that the two-elbow disturbances can
make a 23% shift in the standard discharge coefficient 4. However, the combination of a cone swirler with a two-elbow disturbance causes a change in discharge coefficient up to 46%. It
means that the cone swirler conditioner on its own can cause up to
a 6% change in standard discharge coefficient, and the combination of this device with other disturbances such as block disturbance, swirler disturbance, and two elbow again causes the
change in discharge coefficient within the 1% error bars. In other
words, the cone swirler conditioner can produce a repeatable shift
in discharge coefficient independent of sources of disturbance. On
the other hand, this graph shows that the result of change in C for
using the standard orifice plate falls within the 1% error bars too.
The latter conclusion confirms the accuracy of the standard orifice
metering 1.
3.3 Performance of the Cone Swirler at Low Reynolds
Numbers (Air Rig). After the measurement of the cone swirler on
a disturbed flow at relatively high Reynolds number, which was
achievable for water rigs, this procedure was examined on the air
rig with low Reynolds numbers. In the air rig, low Reynolds numbers up to 20,000 were obtained. The results for the air rig with
different combinations of disturbances and the cone swirler are
shown in Fig. 11.
The graph shows that using a cone swirler conditioner in the
upstream changes the standard discharge coefficient by up to
1.5%. This trend is repeated when a combination of block disturbance and cone swirler is placed upstream. This means that the
Fig. 10 Change in discharge coefficient for two-elbow disturbance with cone swirler
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Fig. 11 Change in discharge coefficient for different disturbances and cone swirler air
rig
cone swirler can attenuate the effect of asymmetric flow with low
Reynolds number. On the other hand, employment of swirler disturbance makes up to 1.5% error on metering, and this change
when using a cone swirler conditioner can be affected by 0.5%
error. This means that the cone swirler cannot compensate the
effect of swirling flow as much as for asymmetric flow.
3.4 Calibration of the Cone Swirler. While the swirler conditioner is used in front of an orifice plate, the standard discharge
coefficient equation 2 cannot be used to calculate a mass flow
rate. So the appropriate discharge coefficient for the combination
of the swirler and an orifice plate was determined by curve fitting.
Figure 12 shows the calibration curve for the combination of
cone swirler and orifice plate. The experimental mass flow rate in
3.6 Source of Errors. A brief review of error sources associated with measurement of the various parameters is considered in
this section. The accuracy of the weighing scale was 0.1% and the
stop watch 0.2%. Also, as already mentioned, the pressure drop
was measured by two kinds of pressure sensors: a pressure transducer and U-tube manometer. The pressure difference between
these two meters was compared, and the difference was about 1.8
cm H2O or 176 Pa in 3 m, which is negligible in metering. On the
other hand, the same procedure was used for measuring the mass
flow rate, and so the inaccuracy was the same for the mass flow
rate. Thus, it can be concluded that the errors in metering are
within those predicted by the Standards.
MAY 2009, Vol. 131 / 051104-7
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Conclusions
The present study has shown that the novel idea of using a
swirler flow conditioner for disturbed flow can reduce the error of
metering caused by disturbances to an acceptable standard level.
To discover the best design of the swirler flow conditioner, a
variety of different shapes was considered and tested in experimental rigs. So far the results show that the cone swirler flow
conditioner can reduce the distortion due to asymmetric velocity
profile on metering for high and low Reynolds numbers. It is
concluded that this kind of swirler conditioner also has a positive
effect on swirling flow for high Reynolds numbers but has less
effect for low Reynolds numbers. The new discharge coefficient
appropriate for the cone swirler flow conditioner can vary from
4% for low Reynolds numbers up to 67% of standard discharge
coefficient for high Reynolds number.
Regarding other factors for desired flow conditioner, the results
show the cone swirler to be positioned about 1.5 pipe diameters
upstream of the orifice plate and its pressure loss to be 2.3 of
dynamic head. These values clearly describe the advantages of the
cone swirler flow conditioner in comparison with other vortex
action flow conditioners.
Also, an important feature is that the standard orifice plate is
vulnerable for metering of dirty flow because of accumulation of
dirt in front of the orifice plate. So the accumulated particles can
alter the accuracy of metering dramatically. Thus using the cone
swirler conditioner not only gives a mass flow metering independent of upstream disturbance but also keeps particles in suspension and prevents the accumulation of particles in front of the
orifice plate.
Nomenclature
d
D
C
p
Re
P
u
C
Cd0
Cd
References
1 BS EN ISO 5167, 2003, Measurement of Fluid Flow by Means of Pressure
Differential Devices Inserted in Circular Cross Section Conduits Running
Full, British Standards Publications.
2 Rogers,
S.,
2003,
Differential
Pressure,
available
on
www.flowcontrolnetwork.com/PastIssues/julaug1999/1.asp.
3 Ouazzane, A. K., and Benhadj, R., 2002, Flow Conditioners Design and Their
Effects in Reducing Flow Metering Errors, Sens. Rev., 223, pp. 223231.
4 Laribi, B., Wauters, P., and Aichouni, M., 2002, Experimental Study of Aerodynamics Behaviour Downstream of the Three Flow Conditioner, ASME
Fluids Engineering Division Summer Meeting, Canada.
5 Reader-Harris, M. J., Hutton, S. P., and Laws, E. M., 1989, Flow Straighteners and Flow Conditioning Devices, Flow Measurement and Instrumentation
Consortium Report No. 7.
6 Miller, R. W., 1996, Flow Measurement Engineering Handbook, 3rd ed.,
McGraw-Hill, New York.
7 National Engineering Laboratory NEL, 1998, Flow Conditioners Performance Review, Guidance Note No. 11.
8 National Engineering Laboratory NEL, 2002, Assessment of the Performance of Flow Conditioner at Elevated Reynolds Number, Guidance Note
No. 29.
9 National Engineering Laboratory NEL, 1999, CFD Techniques Applied to
Differential Pressure Flowmeter Performance, Guidance Note No 20.
10 Frattolillo, A., and Massarotti, N., 2002, Flow Conditioners Efficiency a
Comparison Based on Numerical Approach, Flow Meas. Instrum., 13, pp.
111.
11 Canada Pipeline Accessories, 2004, CPA 50E Flow Conditioner, available
on www.flowmeterdirectory.com/flowmeter_flow_conditioners.html.
12 Laws, E. M., and Ouazzane, A. K., 1995, A Further Investigation Into Flow
Conditioner Design Yielding Compact Installations for Orifice Plate Flow Metering, Flow Meas. Instrum., 63, pp. 187199.
13 2004,
In-Line
Flow
Conditioning
Plate,
available
on
www.sierrainstruments.com/products/pdf/flow-track.pdf.
14 2004, AS-FC Flow Conditioner, available on www.fluidcomponents.com/
Aerospace/A_ProdflowCond_ASFC.htlm.
15 Ouazzane, A. K., and Barigou, M., 1999, A Comparative Study of Two Flow
Conditioners and Their Efficiency to Reduce Asymmetric Swirling Flow Effects on Orifice Meter Performance, Trans. Inst. Chem. Eng., Part A, 77, pp.
747753.
16 Beck, S. B. M., and Mazille, J., 2002, A Study of Pressure Differential Flow
Meter That Is Insensitive to Inlet Conditions, Flow Meas. Instrum., 125-6,
pp. 379384.
17 Ahmadi, A., Beck, S., and Stanway, R., 2003, Effect of Diameter Ratio on the
Swirling Orifice Plate, Seventh International Symposium on Fluid Control,
Flow Measurement and Flow Visualization, Italy.
18 Ahmadi, A., and Beck, S., 2005, Development of the Orifice Plate With a
Cone Swirler Conditioner, Sens. Rev., 251, pp. 6368.
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
2D NavierStokes Simulations of
Microscale Viscous Pump With
Slip Flow
Khaled M. Bataineh1
e-mail: k.bataineh@just.edu.jo
Mohd A. Al-Nimr
Department of Mechanical Engineering,
Jordan University of Science and Technology,
Irbid 22110, Jordan
In this paper we provide numerical solution of the NavierStokes equations coupled with
energy equation for gaseous slip flow in two-dimensional microscale viscous pumps. A
first-order slip boundary condition was applied to all internal solid walls. The objectives
are to study the performance of the pumps and to study the effect of velocity slip on its
performance. Mass flow rate and pump efficiency were calculated for various pump
operation conditions when an external pressure load is applied at the pump exit plane.
Geometric parameters were held fixed in this work. Microviscous pump performance was
studied in detail for several values of the Reynolds number, pressure load, eccentricity,
and slip factors. Our numerical results for no-slip were compared with previously published experimental and numerical data and were found to be in very good agreement.
Slip values and eccentricity were found to be major parameters that affect the performance of pump. Pump head decreases with increasing slip factors. Maximum pump
efficiency increases with increasing slip factor up to Kn approaching 0.1. However, the
maximum value of pump efficiency is found to experience a steep degradation for Kn
approaching 0.1. The values of moment coefficient always decrease as both slip factor
and distance of the rotor from the lower wall increase. Also, as slip factors and distance
of the rotor from the lower wall increase, less net flow rate is predicted. For a given fixed
driving force at the rotor surface, there is an optimum value for the behavior of pump
efficiency with distance of the rotor from the lower wall. Future research should be
conducted to modify the current design to make this concept work for higher Knudsen
numbers. DOI: 10.1115/1.3112390
Keywords: microfluidics, viscous micropump, slip flow, low Reynolds number, CFD
Introduction
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Governing Equation
x
,
D
y =
y
,
D
H =
H
,
D
Lu =
Lu
,
D
Ld =
Ld
,
D
d
D
1
where Ld is the downstream distance, Lu is the upper stream distance, H is the channel height, and is the cylinder eccentricity.
We define the dimensionless x- and y-component of fluid velocity as
u =
u
D/2
v =
v
D/2
= r o
where o is the density at the exit of the channel, which is assumed to be constant, without loss of generality, we assume r to
be equal to unity.
The nondimensional pressure rise pump load is defined as:
p =
Pout Pin
2/D2
where Pout is the exit pressure, Pin is the inlet pressure, is the
density of fluid, is the rotational speed of the shaft, and is the
kinetic viscosity of the fluid.
We define the drag, lift, and moment coefficients as follows:
CD =
CL =
CM =
FD
1
oU 2D
2
FL
1
oU 2D
2
M
1
oU 2D 2
2
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
FD =
Rp cos d
Rr sin d
Table 1 Demonstrates the mesh convergence of dimensionless average flow velocity u, Re= 1, H = 1.5, and = 0.025
Rp sin d
Rr cos d
10
No. of elements
11
639
1563
6413
22,352
0.1001
0.10368
0.10514
0.1058
ui ui+1
100%
ui
M=
R 2 rd
12
Re2V V = P + Re 2V
13
where Re= D2 / 2 is the Reynolds number based on rotor angular velocity and shaft diameter.
A gas flow is classified to be in the slip regime if the Knudsen
number Kn is 0.001 Kn 0.1. Here the Knudsen number is defined as Kn= / H, where is the mean free path of gas molecules. For macrofluidic devices, Kn is very small, and no-slip
boundary condition at solid boundaries is a valid assumption. In
micromechanics of gases, Kn is not small and there is a slip at
solid boundaries 1517. The traditionally used no-slip boundary
conditions for velocity are no longer valid in the micromechanics
of gases. In the slip regime, the gas velocity at a solid surface
differs from the velocity at which the wall moves. Following
Maxwell 18, it has been proposed by Schaaf and Chambre 19
that at a solid surface in isothermal slip flow, the dimensionless
slip velocity is defined as
vs
vg
2 vg
+
Kn
14
15
FL =
N/A
3.466
1.40313
0.6076
2
U
2
Kn
U Uc
g
Vg V ng = Vw
16
17
k BT
2v2p
18
Numerical Modeling
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
0
7.5
15
30
u Present work
u 6
0.1058
0.075
0.042
0.0214
0.106
0.0736
0.043
0.021
Difference
upresent
uSharatchandra
100%
uSharatchandra
%
0.19
1.9
2.3
1.9
1
H
uydy
19
5.1 Effects of Slip Factor and Pump Load. The pump load
pump backpressure is modeled by increasing the outlet pressure
of the pump to simulate the pressure increase needed from the
pump to overcome the imposed external pressure. The effect of
slip was simulated by changing the values of slip factor. We investigated four different values of slip factor, namely, S = 0,
0.0027, 0.027, and 0.27. Figure 3 shows the change in the flow
rate at different pump loads and slip factors. It is clear that the
pump flow rate decreases linearly with increasing pump load. It is
also clear that pumping flow rate decreases with increasing slip.
When the external imposed pressure exceeds the maximum load
the pump can deliver, a backflow occurs and the average velocity
becomes negative. The maximum load that pumps can deliver
decreases with increasing slip. For example, the maximum pumping load is around 25 for no-slip. This value decreases dramatically to a value around 4 for Kn approaching 0.1 corresponds to
S = 0.27. Figure 3 shows that for S = 0.0027 corresponds to Kn
= 0.002, the deviation from no-slip is small. This behavior agrees
very well with the accepted classification of flow regime, that is,
for Kn greater than 0.001, the no-slip is a valid condition. From
the definition of S, S = 0.27 corresponds to Kn approaching 0.1 in
our simulations. For Kn approaching 0.1, the pump operating conditions degraded significantly. The maximum backpressure at
zero flow rate is only 0.0252 Pa. The maximum flow rate ob-
Fig. 4 Total drag coefficient viscous and pressure as a function of pump load for different slip factors = 0.025, H = 1.5, and
Re= 1
20
where the input mechanical energy is M . M is the torque supplied to the rotor and is the angular velocity. The net energy of
the flow is only in the form of pressure increase at the pump
outlet, since the kinetic energy at the inlet and exit are equal and
there is no elevation change. Hence, micropump efficiency can be
obtained by
p
m
=
M
21
Fig. 5 Moment coefficient as a function of pump load for different slip factors = 0.025, H = 1.5, and Re= 1
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
upH
C M Re2
22
Simulations were carried out to determine the micropump efficiency and the effect of the flow parameters on the performance of
the pump. Figure 6 shows how the efficiency varies as a function of pump load and slip factors. The pump efficiency is high
when the flow rate and pump load are high and the coefficient of
moment is low. Since for a fixed slip factor, the coefficients of
moment do not change considerably, the dominant terms are the
pump load and flow rate. Since the flow rate is inversely proportional to the pump load, there should be an optimum operating
condition of the pump. Figure 6 shows that the optimum operating
condition lies in the middle the pump loads. For example, for
no-slip, the pump delivers flow when P varies from 0 to 25 see
Fig. 3. Figure 6 shows that the maximum value of efficiency
corresponding to pump load is equal to half of the maximum
backpressure zero flow rates. The same is true for different values of slip factor. Generally, the effect of increasing the slip factor
is the increase in pump efficiency and the shift of the optimal
point to the left. For given values of H, P, and Re, pump
efficiency reduces to the ratio between u and C M . Increasing slip
factor reduces both u and C M . Numerical results predict that C M
decreases more than u with increasing slip factor. However, when
Kn approaches 0.1, average velocity experiences a steeper reduction, while moment coefficient still decreases smoothly. Figure 6
shows that pump efficiency has been greatly degraded for S
= 0.27 compared with S = 0.027 or smaller. This agrees well with
Ref. 6. For example, for Kn approaching 0.1, maximum efficiency is only about 1.3%. Maximum efficiency for no-slip is
about 2.2%, which matches previously published results.
function of Reynolds number and slip factors. The flow rate increases linearly with increasing Reynolds number. Also, flow rate
decreases with increasing slip factor.
Figure 8 shows that pump efficiency stayed constant as a function of the Reynolds number. This was due to the fact that efficiency is inversely proportional to the product of C M and Re2.
Figure 9 shows that C M and CD decrease with increasing Reynolds number, so the product stayed almost constant.
5.3 Effect of Rotor Eccentricity. Simulations were carried
out to investigate the effect of rotor eccentricity on pump performance for several values of slip factors. External pressure was
fixed to 1. Reynolds number was fixed to unity. Figure 10 shows
flow rate as a function of rotor eccentricity for different slip factors. As S and increase, less net flow rate is predicted. Although
higher values of S creates more forward flow in the upper gap and
more backward flow in the lower gap, but the net effect will be
less net flow as S increases. Also, as increases, the width of the
Fig. 9 Moment coefficient as a function of pump load for different slip factors = .025, H = 1.5, and Re= 1
Fig. 10 Flow rate as a function of rotor eccentricity for different slip factors Re= 1, H = 1.5, and P = 1
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
lower gap increases and that of the upper gap decreases. This
causes a reduction in the amount of the forward flow in the upper
gap and an increase in the amount of the backward flow in the
lower gap. The net effect will be a reduction in Q with increasing
. Our results agreed well with previously published data that flow
rate increases with decreasing .
Figure 11 shows the moment coefficient as a function of rotor
eccentricity for different slip factors. The values of C M always
decrease as both S and increase. Increasing S implies that the
slipping effects at the rotor surface increases and this yields less
velocity gradients and as a result, less shear forces and C M . In the
same manner, as increases, the slipping effects will decrease due
to the less velocity gradients near the rotor surface. As a result,
less C M values are predicted. However, for S = 0.27, the moment
coefficient reduces slightly with increasing .
Figure 12 shows efficiency variation with S and . For small
as S increases the efficiency increases. For a given fixed driving
force at the rotor surface and as S increases, the flow in both the
upper and lower gaps will be enhanced by the slipping effects due
to the less wall resistance against the flow. This implies that both
the forward flow in the upper gap and backward flow in the lower
gap will increase as S increases. The flow in the upper wider gap
is subject to less slipping effects due to its less velocity gradients,
as compared with the flow in the lower gap. But since the upper
gap has a larger area the net effect will be enhancing the forward
flow as S increases and hence improving pump efficiency. At large
the lower gap has a significant area and the effect of will be to
decrease regardless of the slipping effects. Within this limit of
relatively larger , the increase in will increase the area of the
backward flow and decrease the area of the forward flow. The net
effect will be a reduction in the net flow as increases and this
causes a reduction in regardless of the values of S, because at
large values of the slipping effects become insignificant.
Tracing the behavior of a single curve one may observe that
there is an optimum value for the behavior of with . There are
Fig. 12 Efficiency as a function of rotor eccentricity for different slip factors Re= 1, H = 1.5, and P = 1
Conclusion
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Nomenclature
CD
CL
CM
d
D
Kn
H
kB
L
Ld
Lu
M
Pin
Pout
P
Re
S
T
u
u
U
vs
x, y
v
V
vg
drag coefficient
lift coefficient
moment coefficient
rotor distance from lower wall m
rotor diameter m
Knudsen number
channel height m
Boltzmann constant
channel length m
down stream distance m
upperstream distance m
moment N / m2
inlet pressure Pa
outlet pressure Pa
pressure Pa
Reynolds number
slip factor
torque N m
x-component velocity m/s
mean velocity m/s
rotor surface velocity m/s
tangential velocity of the solid surface m/s
Cartesian coordinates m
y-component velocity m/s
velocity vector m/s
velocity of the gas m/s
cylinder eccentricity
pump efficiency
mean free path m
kinematic viscosity m2 / s
density at channel exit kg/ m3
density kg/ m3
density
tangential-momentum accommodation
coefficient
References
1 Judy, J. W., 2001, Fabrication, Design and Applications, Smart Mater.
Struct., 10, pp. 11151134.
2 Ziaie, B., Baldi, A., Lei, M., Gu, Y., and Siegel, R. A., 2004, Hard and Soft
Micromachining for BioMEMS: Review of Techniques and Examples of Applications in Microfluidics and Drug Delivery, Adv. Drug Delivery Rev., 56,
pp. 145172.
3 Gad-el-Hak, M., 1999, The Fluids Mechanics of MicrodevicesThe Freeman Scholar Lecture, ASME J. Fluids Eng., 121, pp. 533.
4 Sen, M., Wajerski, D., and Gad-el-Hak, M., 1996, A Novel Pump for MEMS
Applications, ASME J. Fluids Eng., 118, pp. 624627.
5 Stemme, E., and Stemme, G., 1993, Valveless Diffuser/Nozzle-Based Fluid
Pump, Sensors and Actuators, Physica, 39, pp. 159167.
6 Sharatchandra, M. C., Sen, M., and Gad-el-Hak, M., 1997, NavierStokes
Simulation of a Novel Viscous Pump, ASME Trans. J. Fluids Eng., 119, pp.
372382.
7 Bart, S. F., Tavrow, L. S., Mehregany, M., and Lang, J. H., 1990, Microfabricated Electrohydrodynamic Pumps, Sens. Actuators, A, 2123, pp. 193
197.
8 Odell, G. M., and Kovasznay, L. S. G., 1971, A New Type of Water Channel
With Density Stratification, J. Fluid Mech., 50, pp. 535543.
9 Sen, M., Wajerski, D., and Gad-el-Hak, M., 1996, A Novel Pump for MEMS
Applications, ASME J. Fluids Eng., 118, pp. 624627.
10 Sharatchandra, M. C., Sen, M., and Gad-el-Hak, M., 1998, Thermal Aspects
of a Novel Viscous Pump, ASME J. Heat Transfer, 120, pp. 99107.
11 Abdelgawad, M., Hassan, I., and Esmail, N., 2004, Transient Behavior of the
Viscous Micropump, Microscale Thermophys. Eng., 8, pp. 361381.
12 Abdelgawad, M., Hassan, I., Esmail, N., and Phutthavong, P., 2005, Numerical Investigation of Multistage Viscous Micropump Configurations, ASME J.
Fluids Eng., 127, pp. 734742.
13 da Silva, A. K., Kobayashi, M. H., and Coimbra, C. F. M., 2007, Optimal
Theoretical Design of 2-D Microscale Viscous Pumps for Maximum Mass
Flow Rate and Minimum Power Consumption, Int. J. Heat Fluid Flow, 28,
pp. 526536.
14 Matthews, M. T., and Hill, J. M., 2008, Lubrication Analysis of the Viscous
Micro/Nano Pump With Slip, Microfluid. Nanofluid., 45, pp. 439449.
15 Fuhr, G., Hagedorn, R., Mller, T., Benecke, W., and Wagner, B., 1992, Microfabricated Electrohydrodynamic EHD Pumps for Liquids of Higher Conductivity, J. Microelectromech. Syst., 1, pp. 141146.
16 Richter, A., Plettner, A., Hofmann, K. A., and Sandmaier, H., 1991, A Micromachined Electrohydrodynamic EHD Pump, Sens. Actuators, A, 29, pp.
159168.
17 Harrison, D. J., Manz, A., and Glavina, P. G., 1991, Electroosmotic Pumping
within a Chemical Sensor System Integrated on Silicon, Proceedings of the
International Conference on Solid-State Sensors and Actuators Transducers,
pp. 792795.
18 Maxwell, J. C., 1879, On Stresses in Rarified Gases Arising From Inequalities
of Temperature, Philos. Trans. R. Soc. London, 170, 231256.
19 Schaaf, S. A., and Chambre, P. L., 1961, Flow of Rarefied Gases, Princeton
University Press, Princeton, NJ.
20 FLUENT 6.0 User Guide Manual, 2002, Fluent, Inc., NH.
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Introduction
properly model the flow field an unsteady simulation of the complete rotor needs to be performed. The instabilities that resulted in
the rotor meant it experienced as much as a 2 deg change in the
inlet flow angle near stall over a 510 revolution period 1.
This study aims to more accurately quantify and present the
magnitude and frequency of these instabilities in a transonic rotor.
To do this simultaneous results of a number of sensors are combined to show the distribution of the instabilities over the rotor.
The frequency and distribution of the instabilities are shown at
various operating points each successively closer to stall. The distributions are of interest as the region of greatest magnitude of the
instabilities occurred in the complex flow region containing the
tip-vortex shock interaction and shock boundary-layer interaction.
In addition as stall was approached the low-frequency instabilities
are detectable well upstream and downstream of the rotor.
The detection of these low-frequency instabilities required the
use of high-speed pressure sensors and good signal processing
techniques. It will be shown that the instabilities do not behave in
either a linear fashion toward stall nor do they occur at predictable
frequencies. In addition the magnitudes of the instabilities were
significant, which resulted in observable changes in the bulk flow
properties.
There may be many possible mechanisms for the production of
these low-frequency instabilities. The fact that the present test
case was a rotor-only machined blisk seemed to indicate that they
were inherent in the flow without the need for external excitation.
Further improvement to stall margin for stable operation of future
compressors close to stall will demand better methods to simulate
these low-frequency instabilities and better understanding of their
causes and effects on machine operation.
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Experimental Apparatus
Experimental Program
70
90
95
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
F was at the full open throttle position of the test rig, point E
was at peak efficiency, and point D was where low-frequency
instabilities first appeared. These points are used for reference
throughout the rest of the paper.
Postprocessing
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
follows. The desired cutoff frequency f c must be nondimensionalized by the sampling frequency
fc =
cuttoff frequency
sampling frequency
The roll-off bandwidth BW was defined as the desired frequency width where the signal dropped from 99% to 1% in
strength. The constant BW was defined as follows
BW =
roll-off width
sampling frequency
2
2
:1:
BW BW
sin2 f cM
2 f c M
Due to the filter length being finite a truncation error occurs that
requires correction. To achieve this, a Blackman window is used
wBlack and results in near unity gain in the band-pass region
2
4
+ 0.08 cos
M
M
Black
To create a high-pass filter, all the elements of hBlack were multiplied by 1 and unity added to the central sample. A low-pass,
h1, and high-pass, h2, filter were then combined to form a bandpass filter. Multiple band-pass filters can be added together but
only one was used here. With the desired digital filters constructed
the steps outlined earlier can be used to filter the sampled signal S
to obtain the filtered signal S f
S f = IFFTFFTS FFTh1 FFTh2
4.2 Frequency Domain. Observation of the signals in the frequency domain was also useful. In order to better isolate the lowfrequency signals the signals were transformed into the frequency
domain using FFTs as these were computationally cheap and easily implemented. A spectral analysis of the frequency data showed
the magnitude and frequency of the sampled pressures over the
rotor. The long data samples and high sampling rate ensured minimal aliasing and a good signal to noise ratio. To ensure that the
correct magnitude of a particular frequency was captured the
sampled signals were padded with zeros, usually eight times the
length of the sample. Practically this was done by performing a
longer FFT than the number of points in the sample. If this very
simple method was not used the magnitude would often be
cropped especially at lower frequencies 10, which were of special interest in this research.
051106-4 / Vol. 131, MAY 2009
Results
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Mean inlet
flow angle
deg
rms variation
deg
Peak-to-peak variation
deg
95% A
90% A
70% B
70.6
71.5
75.1
0.351
0.341
0.269
2.04
2.21
1.65
Mean pressure
Raw
peak to peak
Filtered
peak to peak
Ratio
10
1
4
6
1.00
0.92
1.25
1.30
0.0502
0.5649
0.3281
0.1269
0.0048
0.0200
0.0359
0.0234
0.092
0.035
0.109
0.184
Sensor No.
Mean pressure
Raw
peak to peak
Filtered
peak to peak
Ratio
Sensor No.
Mean pressure
Raw
peak to peak
Filtered
peak to peak
Ratio
10
1
4
6
1.00
0.90
1.29
1.35
0.094
0.6069
0.3767
0.1453
0.0096
0.0215
0.0406
0.0351
0.102
0.035
0.107
0.241
10
1
4
6
1.00
0.97
1.14
1.17
0.0335
0.1426
0.1826
0.1029
0.0022
0.0074
0.0193
0.0100
0.065
0.052
0.106
0.097
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
10
Blade passing
frequency
10
Magnitude
10
10
10
95% Speed
Kulite 4
12
10
.1
10
Frequency (Normalized to Rotor)
100
x 10
Magnitude
2.5
95% Speed
Kulite 4
1.5
0.5
0.3
0.4
0.5
0.6
0.7
Frequency (Normalized to Rotor)
0.8
0.9
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Discussion
The ability to observe the collective behavior of the lowfrequency instabilities allowed some insight into their possible
causes and the prestall characteristics of some transonic machines.
It is important to recall that the periodic assumption of the flow in
each passage being the same and steady relative to the blade does
not always hold when operating close to stall 1,5. Typical nearstall flow of a transonic machine contains a normal detached
shock ahead of the blades with a tip-vortex that intersects the
shock, which distorts it as shown in Fig. 5. The shock impinges on
the suction surface of the blade ahead of it.
6.1 Protruding Instabilities. The so called protruding instabilities indicated in the previous figures were most likely due to
the movement of the normal shock relative to the blade row as the
shock clearly protruded upstream of the blade. At peak efficiency
the shock in a transonic rotor is attached to the leading edge and
thus geometrically fixed to the rotor. A detached normal shock
would be affected by any change in the passage blockage and a
changing position relative to the blade would have been detected
upstream of the rotor. This movement may be due to multiple
mechanisms as indicated by the more than one protrudinginstability that were sometimes present Fig. 13, A and B. The
tip-vortex may also not be steady near stall and could affect the
normal shock, which of course in turn would affect the following
blades flow.
MAY 2009, Vol. 131 / 051106-9
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
6.2 Peak Instability Region. The largest instabilities that appeared were within the blade passage at sensors 35 at all
speeds. In this region the tip-vortex would have been interacting
with the detached normal shock as mentioned before. This also
corresponded axially to the location of the shock impinging on the
blade suction surface. While the vortex-shock interaction has a
strong influence in the tip region the shock-boundary impingement region occurs for a significant part of the span in a transonic
machine especially at higher Mach number.
6.3 Instability Behavior. The behavior of the instabilities
was generally nonlinear when approaching stall with different frequencies becoming dominant as stall was approached. It is important to note, however, that as long as they remain below a certain
magnitude stall would not occur. The main difference between
95% and 90% speeds was that at 95% speed the two main instabilities coalesced. At 95% speed, one of the instabilities presumably acted as a strong forcing function causing the other to follow
suit Fig. 11. At 90% speed Fig. 13 the two instabilities remained independent.
The exact mechanism that causes the instabilities can only be
hypothesized here. They clearly moved relative to the rotor but
only appeared when the compressor began to approach stall. The
types of flow found in a rotor, such as the tip-vortex, its interaction with the shock, and the detached normal shock would vary
with small disturbances in the external flow field. A variation due
to a small disturbance in one passage would affect the flow in the
following passage akin to a stall cell. This type of behavior could
reinforce itself after a number of trips around the rotor and eventually lead to the larger detectable variations that were observed.
As stall is approached the adverse pressure gradients increase and
flows with adverse pressure gradients such as in diffusers and
rotors tend to amplify flow disturbances.
6.4
6.5 Stall Prediction Improvements. The presence of instabilities, especially at multiple frequencies, makes the exact prediction of stall difficult. As noted it may take a number of revolutions
of the machine for the conditions in the passage that cause a stall
to be present. This may be different for each stall event. When
simulating the flow making using of a periodic boundary assumption for the blade rows this in effect numerically damps out the
low-frequency instabilities. Whether this makes the simulated prediction of stall pessimistic or optimistic depends on the machine.
Ideally full rotor simulations are needed and the simulations will
have to be conducted over many revolutions. If a full annulus,
transient simulation was performed the shocks, tip-vortex, and
boundary layers would be allowed to vary between passages. As
was noted though the actual change in the flow of a single passage
is quite slow, around 14 chord lengths peak-to-trough, which may
leave some scope to using existing single passage steady-state
methods to improve designs. A design change that would indicate
a stall-margin improvement in a steady-state simulation is likely
to lead to a stall-margin improvement in the real machine. However, if the low-frequency instabilities increased in magnitude as
this study seems to indicate the magnitude of the improvement is
likely to be less than predicted.
Conclusions
Acknowledgment
The present study was part of the compressor research program
sponsored by the Propulsion and Power Department of the Naval
Air Warfare Centre, Patuxent River, MD with Ravi Ravindranath
as the technical monitor.
Transactions of the ASME
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
References
1 Gannon, A. J., and Hobson, G. V., 2007, Pre-Stall Modal Instabilities in a
Transonic Compressor Rotor, ISABE, Beijing, China.
2 McDougall, N. M., Cumpsty, N. A., and Hynes, T. P., 1990, Stall Inception in
Axial Compressors, ASME J. Turbomach., 112, pp. 116125.
3 Camp, T. R., and Day, I. J., 1997, A Study of Spike and Modal Stall Phenomena in a Low-Speed Axial Compressor, ASME Turbo, Orlando, FL, Paper No. 97-GT-526.
4 Bergner, J., Kinzel, M., Schiffer, H.-P., and Hah, C., 2006, Short LengthScale Rotating Stall Inception in a Transonic Axial Compressor: Experimental
Investigation, ASME Turbo, Barcelona, Spain, Paper No. GT2006-90209.
5 Hah, C., Bergner, J., and Schiffer, H. P., 2007, Rotating Instability in a Transonic Compressor, ISABE, Beijing, China.
6 Sanger, N. L., 1996, Design of a Low Aspect Ratio Transonic Compressor
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
B. M. Marino
CONICET Researcher
Associate Professor
e-mail: bmarino@exa.unicen.edu.ar
L. P. Thomas
CONICET Researcher
Associate Professor
e-mail: lthomas@exa.unicen.edu.ar
Instituto de Fsica Arroyo Seco,
Universidad Nacional del Centro,
Pinto 399,
B7000GHG Tandil, Argentina
Introduction
Gravity currents are flows driven by horizontal buoyancy differences of a fluid running inside another fluid of slightly different
density, at the top, at the bottom, or at an intermediate level. As
they are present in numerous important natural and industrial
events, the comprehension of the motion of its frontal region is
particularly important for a variety of practical situations. For example, the analysis of gravity currents is relevant for water quality
management in reservoirs as they carry suspended matters and
dissolved solids across the lake, often determining the distribution
of pollutant substances. Analogous environmental flows are
caused by the spreading of the cooling water from a power plant
in a river, the spreading of cool marine air under the warmer air
overlying land which is known as the sea-breeze, avalanches of
airborne snow particles, fiery avalanches and base surges formed
from gases and solids issuing from volcanic eruptions, and the
spread of accidentally released liquid natural gas. The properties
of this diverse range of dynamically related flows have been described and comprehensively reviewed by Simpson 1.
In particular, estuaries also exhibit a variety of phenomena
driven ultimately by the density differences associated with freshwater discharge, which may be understood in terms of gravity
current theory. One characteristic of the estuarine phenomenon is
the tidal intrusion of brackish water formed under low energy
conditions of the estuary and characterized by a range of small
magnitude tides and weak effects of the wind over the flow, as
typically occurs in narrow canals. The penetration of this saltwater wedge may and frequently does contaminate and become
useless the supply of potable water and water for industrial use in
coastal cities upstream, with serious consequences from the ecological and economical points of view. Although typical of estuaries with constricted connections to the sea, these tidal intrusion
fronts have been observed in reservoirs and lakes where cold turbid river inflow may plunge below the less dense ambient water
2.
The main physic phenomena underlying in many natural situations can be analyzed by means of a simple experimental system
constituted by a channel that is temporarily divided into two sections by a thin vertical barrier. Fresh water fills one section and
Contributed by the Fluids Engineering Division of ASME for publication in the
JOURNAL OF FLUIDS ENGINEERING. Manuscript received May 13, 2008; final manuscript
received January 21, 2009; published online April 1, 2009. Assoc. Editor: James A.
Liburdy.
salt water the other, and the levels are made equal. As soon as the
barrier is raised, the dense fluid starts to collapse and countercurrents begin to flow in opposite directions, as illustrated in Fig. 1.
However, the geometry for inflow channels to lakes, reservoirs,
and impoundments uses to play a non-negligible role 3. Most of
the experimental works related to these situations have been conducted in channels with a constant width and sloping bottom see
Refs. 4,5 and references therein and few researchers have dealt
with a varying-width channel or diffusers e.g., Refs. 610. It is
well known that the density underflow and plunge line in a diverging channel may have different regimes, and that the flow field
strongly depends on several parameters: inflow Froude number,
inflow aspect ratio width/depth, the divergent angle, and the bed
slope. In addition, a water course may expand its width and depth,
the flow may have a significant momentum, turbulence may be
important, and so on. However, as most studies are conducted in
idealized geometric setups where effects of individual parameters
are analyzed separately e.g., Ref. 9, it is not possible to generalize the available experimental results to fit all cases, especially
for field applications. Supplementing the existing experimental
results and providing insight into processes that are too difficult to
measure in the field or in a laboratory, computational analysis was
also performed see, for example, Ref. 11.
We are particularly interested in the influence of the transversal
variations of the properties of many currents in estuaries, rivers,
and natural or man-made channels. Even though the influence of
the factors mentioned above deserves additional separated investigations, some indication of their effects might be inferred from
previous studies. For example, Britter and Linden 4, and later
Monaghan 12, found that the front velocity parallel to the slope
is constant for angles less than 45 deg on rectangular cross-section
channels; similarly such an unimportant effect might be expected
for nonrectangular cross-section channels. Fronts in river flows
with significant momentum may be analyzed on a moving frame
of reference but taking into account the change in the stress of the
channel contours. Britter and Simpson 13 studied the influence
of the bottom stress on gravity currents generated in rectangular
cross-section channels in the laboratory, and Zhou 14 in estuaries. They found that the bottom stress may dramatically change
the height profile of the dense current, so the influence of the
momentum of the flow should be carefully analyzed for nonrectangular cross-section shapes as well.
Although in most previous studies, authors assumed that the
properties of the flow are independent of the cross-coordinate as
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
z
Initial position of the gate
u2
h2
h1
u1
h0
x
Theoretical Model
gh
bz
bz
for
y0
for
y0
where z and y are the vertical and transversal coordinates, respectively, and and b are constants. In particular, b = w / h0 is determined by the width w of the channel and the height h0 of the
denser fluid at the initial time. The cross-section area s that is
occupied by the denser fluid up to a height h is
s=
bz
bz
dydz =
2bh+1
+1
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Fig. 2 Schematic of the frontal zone of dense a and light b gravity currents
p = pA 2gz
1
2
2 2u 1
Pressure on section CD is
p=
pC 1gz
for
0 z h1
h1 z h0
pdydz +
2u2dydz
2bh0+1
+1
1
AB =
pO + 2u21
2gh0
2
+1
+2
and
u21
gh0
u21
g 1h 0H 1
F21 =
g 1h 0H 1
1 H1+1 2
=
2
H+1
+2 1
1 + H1+1
1
+2
11
13
1/+1
14
and the Froude number given by Eq. 11 for a dense energyconserving gravity current becomes
where g1 = g1 2 / 2.
Equation 11 is found by relating the velocities and heights
corresponding to two sections, AB and CD, separated by a horizontal distance of the order of the head length. If the length of the
head is much smaller than the total current length, planes AB and
CD may be considered to be near to each other so that Eq. 11 is
valid at a front where the fluid velocity and the current height
are discontinuous functions of the horizontal coordinate x, as
usual in the theoretical analyses.
As expected, Eq. 11 with = 0 becomes the well known relationship for the rectangular cross-section channels proposed first
by Benjamin 21 and confirmed later by other authors see, for
instance, Ref. 26,
Journal of Fluids Engineering
= 21 H1+12
F21c = 2 1
10
12
Equating Eqs. 11 and 13, the nontrivial solution for the current
depth results
H1c =
on section CD.
Since there are no external horizontal forces acting on the flow,
the net flux of momentum into the control volume including the
head is zero. Conservation of the horizontal component of the
momentum may then be written as AB = CD. Using the mass
conservation 4, the Froude number F1 for the dense current is
u21
=0
1 1 H12 H1H1
1 + H1
where = 2 / 1.
While Bernoullis equation is valid between points A and O
see Eq. 6 because of the laminar flow existing there, it is not
always valid between O and any point on plane CD since there is
usually energy dissipation in the head. But for currents in which
dissipation can be considered negligible, Bernoullis equation may
be also applied along a streamline between points on planes AB
and CD as, for example, the upper boundary of the channel to find
2bh0+1
+1
+1
pO
2gh0
1 2gh0H1+2
CD =
+1
+2
+2
2
1 2gh0H11 H1+1 + 2u1D
1 H1+1
1
+2
15
16
p = pO 2 1u22 + 1gh0 z
and
p=
17
0 z h1
pO + 2gh0 z
h1 z h0
for
18
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
F22 =
u22
g 2h 0H 2
2 1 1 H2+21 H2+1
+ 2
2 1 + H2+1H2
19
u22
gh0
=0
= 1
2 H21 H2H2
1 + H2
20
21
Equating Eqs. 19 and 21, the height H2c for an energyconserving light gravity current is found to be the solution of the
algebraic equation
+2
+1
+1
1 H1S
+ 21 H1SH1S
2 H1S
=0
22
where H1S = 1 H2c. Thus, the Froude number for a light energyconserving gravity current is
F22c =
2 1 1 H2c+21 H2c+1
+ 2
2 1 + H2c+1H2c
23
Experimental Results
Fig. 3 Experimental setup to generate gravity flows in a channel of concave a, triangular b, and convex c cross-section
shapes
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
1.5
1.0
=2
=1
=0
.5
=
0
1.2
0.8
H2
0.4
0.6
0.4
0.8
1.2
1.6
From the density distribution, the front velocity and the height
of the opposite flows may be estimated. A few seconds after the
release, the currents enter into the constant velocity or slumping
phase that is maintained practically until the fronts reach the corresponding end-walls. Two layers with the most advanced point
separated from the bottom/top rigid contour are quickly developed. Then the height of each fluid layer, that is, h1 and h2 marked
in Fig. 1, was measured almost at the end of the run and at a
density relative difference of 90% with respect to the initial value.
Other criteria tested do not modify the results significantly more
details can be found in Ref. 25. The values of the front velocity
and current height lead to Reynolds numbers between 7000 and
35,000. Due to Reynolds number independence for density currents moving along boundaries when Re 103 27, laboratory
flows reproduced well many of the problems to which the results
of this paper might be directed. It is found that each set of experiments that is, different relative densities for the same crosssection shape gives similar results when dimensionless variables
H, u / gh01/2, and F are determined; hence they are averaged in
the following.
It is worthwhile noting that the reduced gravity g enters into
Eqs. 1, 11, and 19 through the product gh. That is to say, the
global results will not be strongly affected by mixing that reduces
g but increases h or viceversa, in approximately the same extent
as shown by Marino et al. 25. Therefore, the model is expected
to provide the correct scaling laws even if a strong mixing is
present. So, the Froude numbers F1 and F2 at the front should not
be modified significantly by the particular values of the Richardson numbers after the head.
Figure 5 shows the variation with of the dimensionless thicknesses of the currents obtained from Eqs. 14 and 22. In addition, symbols and vertical bars indicate the average values and the
standard deviations, respectively, calculated from the experimental data for different cross-section shapes. Theoretical and experimental results agree well. Note that H1 + H2 = 1 only for rectangular cross-section = 0; in such a case, both currents occupy half
1
the channel depth H1c = H2c = 2 as well known 1. For a nonrectangular cross-section channel H1c 1 / 2 H2c being greater the
difference between the thicknesses for increasing , and H1 + H2
1. A significant difference between the thicknesses obtained
with open and closed channels was not detected.
Journal of Fluids Engineering
0.0
0.0
2.0
0.2
=
=1 0.5
0.4
F2
0.3
0.2
0.0
0.0
F1
0.9
F1(H1) , F2(H2)
0.6
H1
0.6
0.8
1.0
H1 , H2
Fig. 6 Froude numbers of the dense F1 and light F2 currents
as function of the respective heights H1 and H2 for different .
The energy-conserving solutions F1cH1c and F2cH2c given by
Eqs. 15 and 23, respectively, are plotted by thin solid lines.
Conclusions
The values of the Froude number obtained at the front of gravity currents developed in uniform nonrectangular cross-section
channels are studied as a function of a simplified cross-section
shape. Mass and momentum balances in the frontal region of both
dense and light currents running on the bottom and along the
upper boundary, respectively, are considered. It is shown that the
usual relationships obtained for rectangular cross-sections have to
be modified in order to include the effects of the cross-section
shape on the flow. For example, for triangular cross-section
MAY 2009, Vol. 131 / 051201-5
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Acknowledgment
This work was supported by CONICET and ANPCyT Argentina under Grant No. PIP 5893 and Grant Nos. PICT 34088/05
and 1185/06, respectively.
Nomenclature
F
g
g
h
H
p
Re
s
u
x, y, z
w
, b
Froude number
acceleration due to gravity
reduced gravity
height of the current behind the head
dimensionless thickness of the current
hydrostatic pressure
Reynolds number
area of the channel cross-section occupied by
the denser fluid
frontal zone velocity
longitudinal, transversal, and vertical
coordinates
width of the laboratory channel
parameters associated with the shape of the
channel cross-section
relationship between light and dense fluid
densities
density
kinematic viscosity
Subscripts
0
1
2
AD, CD
c
O
initial value
dense fluid
light fluid
vertical planes limiting the control volume
energy-conserving current
stagnation point
References
1 Simpson, J. E., 1997, Gravity Currents: In the Environment and the Laboratory, Cambridge University, Cambridge, p. 258.
2 Imberger, J., and Hamblin, P. F., 1982, Dynamics of Lakes, Reservoirs and
Cooling Pounds, Annu. Rev. Fluid Mech., 14, pp. 153187.
3 Farrell, G., and Stefan, H., 1988, Mathematical Modeling of Plunging Reservoir Flows, J. Hydraul. Res., 26, pp. 525537.
4 Britter, R. E., and Linden, P. F., 1980, The Motion of the Front of a Gravity
Current Travelling Down an Incline, J. Fluid Mech., 99, pp. 531543.
5 Alavian, V., Jirka, G. H., Denton, R. A., Johnson, M. C., and Stefan, H. G.,
1992, Density Currents Entering Lakes and Reservoirs, J. Hydrol. Eng.,
118, pp. 14641489.
6 Akiyama, J., and Stefan, G., 1987, Onset of Underflow in Slightly Diverging
Channels, J. Hydrol. Eng., 113, pp. 825844.
7 Johnson, T. R., Farell, G. J., Ellis, C. R., and Stefan, H. G., 1987, Negatively
Buoyant Flow in a Diverging Channel. I: Flow Regimes, J. Hydrol. Eng.,
113, pp. 716730.
8 Johnson, T. R., Ellis, C. R., Farell, G. J., and Stefan, H. G., 1987, Negatively
Buoyant Flow in a Diverging Channel. II: 3-d Flow Field Descriptions, J.
Hydrol. Eng., 113, pp. 731742.
9 Johnson, T., Ellis, C., and Stefan, H., 1989, Negatively Buoyant Flow in
Diverging Channel. IV: Entrainment and Dilution, J. Hydrol. Eng., 115, pp.
437456.
10 Stefan, H., and Johnson, T., 1989, Negatively Buoyant Flow in Diverging
Channel. III: Onset of Underflow, J. Hydrol. Eng., 115, pp. 423436.
11 Bournet, P. E., Dartus, D., Tassin, B., and Vinon-Leite, B., 1999, Numerical
Investigation of Plunging Density Current, J. Hydrol. Eng., 125, pp. 584
594.
12 Monaghan, J. J., 2007, Gravity Current Interaction With Interfaces, Annu.
Rev. Fluid Mech., 39, pp. 245261.
13 Britter, R. E., and Simpson, J. E., 1978, Experiments on the Dynamics of a
Gravity Current, J. Fluid Mech., 88, pp. 223240.
14 Zhou, M., 1998, Influence of the Bottom Stress on the Two-Layer Flow
Induced by Gravity Currents in Estuaries, Estuarine Coastal Shelf Sci., 46,
pp. 811825.
15 Nunes, R. A., and Simpson, J. H., 1985, Axial Convergence in a Well Mixed
Estuary, Estuarine Coastal Shelf Sci., 20, pp. 637649.
16 Wong, K. C., 1994, On the Nature of Transverse Variability in a Coastal Plain
Estuary, J. Geophys. Res., 99, pp. 209222.
17 Wong, K. C., and Munchow, A., 1995, Buoyancy Forced Interaction Between
Estuary and Inner Shelf-Observation, Cont. Shelf Res., 15, pp. 5988.
18 Valle-Levinson, A., and Lwiza, K. M. M., 1995, The Effects of Channels and
Shoals on Exchange Between the Chesapeake Bay and the Adjacent Ocean, J.
Geophys. Res., 100, pp. 1855118563.
19 Engqvist, A., and Hogg, A., 2004, Unidirectional Stratified Flow Through a
Non-Rectangular Channel, J. Fluid Mech., 509, pp. 8392.
20 Thomas, L. P., and Marino, B. M., 2004, Lock-Exchange Flows in NonRectangular Cross-Section Channels, ASME J. Fluids Eng., 126, pp. 290
292.
21 Benjamin, T. B., 1968, Gravity Currents and Related Phenomena, J. Fluid
Mech., 31, pp. 209248.
22 von Karman, T., 1940, The Engineer Grapples With Non-Linear Problems,
Bull. Am. Math. Soc., 46, pp. 615683.
23 Lowe, R. J., Rottman, J. W., and Linden, P. F., 2002, A Laboratory Study of
the Velocity Structure in an Intrusive Gravity Current, J. Fluid Mech., 456,
pp. 3348.
24 Birman, V. K., Martin, J. E., and Meiburg, E., 2005, The Non-Boussinesq
Lock-Exchange Problem. Part 2. High Resolution Simulation, J. Fluid Mech.,
537, pp. 125144.
25 Marino, B. M., Thomas, L. P., and Linden, P. F., 2005, The Front Condition
for Gravity Currents, J. Fluid Mech., 536, pp. 4978.
26 Shin, J. O., Dalziel, S. B., and Linden, P. F., 2004, Gravity Currents Produced
by Lock Exchange, J. Fluid Mech., 521, pp. 134.
27 Simpson, J. E., and Britter, R. E., 1979, The Dynamics of the Head of a
Gravity Current Advancing Over a Horizontal Surface, J. Fluid Mech., 94,
pp. 477495.
28 Hartel, C., Meiburg, E., and Necker, F., 2000, Analysis and Direct Numerical
Simulation of the Flow at a Gravity-Current Head. Part 1. Flow Topology and
Front Speed for Slip and No-Slip Boundaries, J. Fluid Mech., 418, pp. 189
212.
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
This study focuses on the detection and characterization of vortices in low Reynolds
number separated flow over the elliptical leading edge of a low aspect ratio, flat plate
wing. Velocity fields were obtained using the time-resolved particle image velocimetry.
Experiments were performed on a wing with aspect ratio of 0.5 for velocities of 1.1 m/s,
2.0 m/s, and 5.0 m/s corresponding to chord length Reynolds numbers of 1.47 104,
2.67 104, and 6.67 104, respectively, and angles of attack of 14 deg, 16 deg, 18 deg,
and 20 deg. A local swirl calculation was used on proper orthogonal decomposition
filtered data for vortex identification and corresponding vortex centers were tracked to
determine convective velocities. The swirl function was also analyzed for its temporal
frequency response at several discrete points in both the shear layer and in the separated
recirculation region. A peak frequency was detected in the shear layer with a corresponding Strouhal number of approximately 3.4 based on the flow direction projected length
scale. The Strouhal number increases with both angle of attack and Reynolds number.
The shear layer convective length scale, based on the vortex convection velocity, is found
to be consistent with the mean separation distance between vortices within the shear
layer. This length scale decreases with increasing Rec. DOI: 10.1115/1.3112385
Introduction
Low and ultralow Reynolds number aerodynamics are of increasing interest in a number of application areas, such as microair vehicles, autonomous vehicles, as well as animal and insect
flight. Low Reynolds numbers based on the freestream velocity
and chord length are typically characterized as less than 106, while
ultralow is typically less than 103. Of concern in the design of
small aircraft is the ability to maintain acceptable lift characteristics over a wide range of angles of attack and to reduce the susceptibility for unstable operation. McCullough and Gault 1 discussed the generally accepted three main types of airfoil stall:
trailing edge, leading edge, and thin airfoil stall. Flow over thin
airfoils at high angles of attack displays attributes of stall where
by the flow separates at the leading edge with a re-attachment
point that moves downstream as the angle of attack increases.
McCullough and Gault likened this to the separation of flow past
a sharp edge because at sufficiently high angles of attack the stagnation point moves below the leading edge.
In this present study, a low aspect ratio, thin wing at high angles
of attack results in leading edge separation with a recirculation
bubble extending over a significant portion of the wing. The separation results in a strong shear layer that yields to a Kelvin
Helmholtz instability. Consequently, as the disturbances grow into
vortical structures along the shear layer, they are convected downstream. Figure 1 illustrates the flow characteristics for the current
study with time averaged velocity vectors and associated streamlines showing the extent of the separation bubble along the centerline for Rec = 1.47 104 and = 20 deg. For all cases studied,
flow reattachment occurs near x / c = 0.3. A flow visualization image of this using a smoke wire technique is shown in Fig. 2, where
flow is from left to right at a chord Reynolds number of approxiContributed by the Fluids Engineering Division of ASME for publication in the
JOURNAL OF FLUIDS ENGINEERING. Manuscript received April 16, 2008; final manuscript received January 26, 2009; published online April 14, 2009. Assoc. Editor:
Juergen Kompenhans. Paper presented at the 2007 ASME International Mechanical
Engineering Congress IMECE2007, Seattle, WA, November 1016, 2007.
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Fig. 1 Mean flow PIV data obtained at the wing centerline. Streamlines
illustrate the extent of the separated region. The light gray region represents an area of no data below the wing. An approximation of the wing
contour is shown with a dotted line.
Fig. 2 Flow visualization of leading edge KelvinHelmholtz instability rollup forming spanwise vortices, Rec = 1.47 104,
= 20 deg.
for wings with blunt trailing edges. They proposed the introduction of a serrated trailing edge that produces streamwise vortices
to disrupted the formation of longitudinal vortex structures.
Yarusevych et al. 10 studied the effect of periodic excitation
of an airfoil as a method of flow control. They investigated matching the excitation frequency with the dominant frequency of separation to improve performance of an airfoil at Rec = 104 and low
angles of attack. Tinar and Cetiner 11 analyzed the self-induced
vibration frequencies of a spring supported airfoil at Reynolds
numbers between 5.9 104 and 14.8 104. Accelerometer data
were paired with particle image velocimetry PIV data to determine natural vibration frequencies and vorticity patterns during
the vibration cycle. Among others, Cheng and Chen 12 performed a numerical study evaluating flow control for the body
forces on a bluff shape and found that dominant frequencies of
vortex shedding corresponded to imposed body force frequencies.
Abernathy 13 investigated flow over an inclined flat plate and
studied the frequency of shed vortices. When analyzing pressure
in the wake of the flat plate, it was seen that the Strouhal number
was nearly constant at 0.17 for angles of attack greater than 35
deg, which corresponds to completely separated flow. This is consistent with the results of Roshko 14 who observed a low range
of Strouhal numbers 0.14 Sth 0.21 for various bluff bodies
and Reynolds numbers.
Burgmann et al. 15 examined the vortex structure of a low
Reynolds number airfoil in water using two-component, timeresolved PIV 2C-TR PIV and scanning PIV. They identified the
development of convex shaped vortex structures ejecting from the
separation bubble behind the leading edge. Derksen and Rimmer
16 employed a vortex cloud model in the simulation of flow
over a low Reynolds number airfoil. This model used free vortices
in the flow that convect over time. The results highlight the importance of such structures in separated flows when considering
body forces such as lift and drag.
1.2 Flow Structure Identification. During the past 2 decades
with the advance of digital PIV, as well as computational fluid
dynamics, large sets of discrete two dimensional velocity fields
have become more readily available for study. While vorticity is a
commonly, and often an appropriately, used measure for detecting
swirl, it is also sensitive to shear and the smallest fluctuations in
the data because it is resolved at the grid level. Several researchers
have set forth alternative methods for quantitatively identifying
coherent structures. Jeong and Hussain 17 identified criteria using the Hessian of the pressure field to identify a vortex core
based on a local pressure minimum. This approach neglects viscous and advective effects in the flow and is based on instantaneous velocity strain rate data. Adrian et al. 18 discussed large
Transactions of the ASME
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Experimental Method
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Hot wire anemometry was used to confirm the TR PIV frequency spectra for a single angle of attack, = 20 deg, at each
Reynolds number. A TSI IFA-300 constant temperature anemometer was used in conjunction with a single axis, TSI 1201-20
model probe. The probe was mounted in the wind tunnel, supported from the downstream direction, and placed in the separation bubble at the centerline approximately 0.25c downstream
from the leading edge. The probe was also placed in the shear
region approximately 0.05c downstream and 0.01c above the
leading edge. Data were obtained using a personal computer PC
with analog capture board and LABVIEW 8.2.
Three freestream velocities were used for this study: 1.1 m/s,
2.0 m/s, and 5.0 m/s, which correspond to chord Reynolds numbers of 1.47 104, 2.67 104, and 6.67 104. Four angles of attack were investigated at each velocity: 14 deg, 16 deg, 18 deg,
and 20 deg, for a total of 12 experimental cases. For each case
images were obtained using an iNanosense high speed digital
camera equipped with an image intensifier. The pixel resolution of
the charge coupled device CCD was 1280 1024 and the field
of view was approximately 55 42 mm2.
Synchronization with the laser pulse was obtained using Dantec
Dynamics FLOWMANAGER software. The time delay between laser
pulses for these data was 50 s for the Rec = 6.67 104 cases and
100 s for the Rec = 1.47 104 and 2.67 104 cases resulting is
an average particle displacement of approximately 6 pixels. The
time-resolved sampling of velocity fields was performed at 500
Hz. The total sample time was 2.0 s due to hardware memory
limitations. This results in a 0.5 Hz frequency resolution when
determining the velocity spectra. It is understood that this may not
be appropriate for analysis of very low frequency events but it
provides adequate results for the higher frequency vortex shedding observed in this study. To help confirm spectral the results,
the TR PIV data were also collected using a 10 s sampling period
for = 20 deg and Rec = 1.32 104, with a sampling rate of 100
Hz. The lower Rec condition was chosen to compare with since it
is the one susceptible to high uncertainty with its lower frequency.
One angle of attack was compared as it has been noted by Abernathy 13 that angle of attack has small influence on normalized
separated flow fluctuations. Overall, the mean velocity vector values and the local standard deviations were all within 2% and 5%,
respectively. Hot wire anemometry, mentioned previously, was
also used to compare low and high frequency events observed in
the TR PIV data and are shown with the results. It should be noted
that the goal of this study is to determine direct spectral characteristics of detected vortical structures rather than velocity components. This data analysis is discussed later.
Postprocessing of the images was done to reduce glare from the
wing surface. This was done by calculating the mean pixel intensity and removing this mean value from each image. The subregion size for each PIV cross-correlation calculation was 32
32 pixels, with an average of 68 particles per subregion. A
50% overlap was used resulting in a vector field of 79 63 grid
points with a spacing of approximately 0.6 mm. The results were
obtained using FLOWMANAGER software from Dantec Dynamic
Denmark using fast Fourier transform FFT cross correlation.
Numerical experiments were carried out with the same seeding
density and particle size as used in the experiments and an adaptive cross-correlation scheme was not found to improve the results
of the swirl detection describe later in a highly curved flow regions 27. The velocity data were filtered using a 3 3 median
filter since large scale flow structures are of interest. No vectors
were removed from the data set. The PIV velocity uncertainties
were investigated in the same wind tunnel, with the same field of
view, while using the same particle generator system and seeding
levels 27. For the same pixel resolution, approximately 6 m
particle diameter, and a 32 32 pixel subregion as used in this
study, the uncertainties were found to be approximately 1% and
velocity gradient bias is small by virtue of estimated particle displacement variation over the subregion domain.
051202-4 / Vol. 131, MAY 2009
Data Analysis
1
AM
PM U M Z
1
dA =
PM U M
AM
sin M dA
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
fx,y,t
d x,ya t
k
Results
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Fig. 4 TR PIV vector field for Rec = 1.47 104 at four angles of attack: a 14
deg, b 16 deg, c 18 deg, and d 20 deg; grayscale plots of are shown
in eh, which were calculated from the velocity fields in ad.
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
are shown in Figs. 7d7f for the = 14 deg cases. The origin
points for tracking are denoted with a shaded star and path lines
are traced through the vortex centers. These results use a POD
filter based on the first 30 modes and a tracking filter time scale,
l = 0.05 or 5% of the convective time scale based on U. By
removing smaller scale fluctuations, the POD filter allows for
more consistent detection and tracking of vortex positions over
time as seen by comparing the vortex position markers circles
between Figs. 7a7c and Figs. 7d7f. This is especially
useful for the highest Rec case.
The centers of the detected swirl regions were determined at all
time steps for each time series. These time series were 2 s long
resulting in 1000 data sets for each flow condition. The centers are
defined based on the geometric mean of contours of = 0.7
similar to those given in Figs. 5f5j. The distance traveled per
sample time was used to calculate the x directional velocity component of these structures. The velocity vector is not strictly only
along the x direction but also the x component is chosen to determine the downstream convection rate for vortical structures generated from leading edge separation. Table 1 lists the calculated
convective velocity, U, normalized by the freestream velocity,
U, for all cases. The standard deviation of U, , normalized by
the freestream velocity is also listed for all cases. Note that no
filtering of the calculated velocities was applied. The convective
velocities show weak variation with angle of attack, less than 5%
Journal of Fluids Engineering
4.2 Spectral Analysis. The time dependent nature of the detected vortices was studied using the transient characteristics of
at all grid location with no threshold applied. The strategy for
evaluating the transient nature of was to examine the regions
where the rms of , , are high as well as the regions where
the rms of the autocorrelation of , , are high. An example
of the nature of the temporal variations of is shown in Fig. 8.
The data shown are taken from a point within the vortex shedding
shear layer at x / c = 0.1, y / c = 0.025. Note that the time span is only
shown for the first 200 ms out of a total of 2000 ms. The autocorrelation of indicates a highly cyclic process with a strong
rms value of the autocorrelation. The associated spectra of
indicates a peak at a Strouhal number, Sth of approximately 2.4.
The distribution of for Rec = 1.47 104, = 20 deg is
shown in Fig. 9a and indicates that fluctuations of are maximum in the shear region. A white line has been added to show the
region of maxima of . Figure 9b shows the distribution of
, which are interpreted as regions of high cyclic characteristics of Two regions of high are identified by enclosed
dashed lines, one just above the region of high rms values of
and the other in the recirculation region. Interestingly, the region
of high rms values of has a relative low cyclic characteristic. In
addition, the high cyclic characteristics within the recirculation
region do not correspond to large amplitude fluctuations since Fig.
9a shows relatively low rms values in the recirculation region.
To help determine the frequency of the cyclic processes within
the shear layer, several experimental results were used. Representative spectra are shown in Fig. 10 for Rec = 1.47 104 and
= 20 deg. These plots are the results averaged over four points
located in the high amplitude cyclic region denoted by the dashed
line above the shear layer shown in Fig. 9b. Using the TR PIV
data at 500 Hz, the average spectrum of the vertical velocity fluctuations, v, is shown in Fig. 10a and the horizontal velocity
fluctuation, u, in Fig. 10b. Both spectra show a peak near Sth
= 3.5, with the v spectrum peak being more dominant. The average spectral for the fluctuations, , is shown in Fig. 10c
revealing a similar peak near Sth = 3.5. These spectra show a broad
low frequency peak in the range of Sth = 0.1 0.3. For all cases the
resolution limit based on the length of the sampling period is
Sth 0.006 0.03, depending on the Reynolds number and angle
of attack. For comparison, hot wire data taken at approximately
the same location in the shear layer were obtained at 10 kHz over
10 s and the resulting spectrum is shown in Fig. 10d, showing a
similar high frequency peak.
The average spectra using six points in the recirculation region
identified in Fig. 9b were obtained using the same methods
stated above for Fig. 10. The average spectra for v and are
shown in Figs. 11a and 11b using the TR PIV data at 500 Hz
for Rec = 1.47 104, = 20 deg. To better illustrate the results, a
MAY 2009, Vol. 131 / 051202-7
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Fig. 6 The POD distribution of energy across modes and angles of attack for a Rec = 1.47 104, b Rec = 2.67 104, and c
Rec = 6.67 104
linear plot is used. The results are shown only up to Sth = 5.0 since
no peaks at higher frequencies were found. Both spectra show
broad maxima near Sth = 0.2. Comparing Figs. 10 and 11 indicates
that the higher frequency peaks in the shear layer occur at approximately an order of magnitude higher value of Sth than any
peaks in the recirculation region. Comparative tests using the hot
wire results and longer time period TR PIV were carried out for a
limited number of cases. Hot wire data obtained at 10 kHz over 10
s are presented in Fig. 11c and TR PIV data sampled at 100 Hz,
over 10 s, are shown in Fig. 11d. These spectra are similar to
those found for the 500 Hz TR PIV data for both v and with
peaks in the same range of Sth.
For all cases of Rec and , the average spectra of were
determined using the four points in the high shear region, which
correspond with regions of high values of . In addition, the
average spectra of using five points in the recirculation region
were also calculated. To illustrate the results, the spectra for
= 20 deg and Rec = 1.47 104 and 2.67 104 are shown in Fig. 12
using the TR PIV data at 500 Hz. Similar to the results shown in
Fig. 10 for the high shear region, Figs. 12a and 12b show the
spectra with peaks occurring at the high frequencies, with broad
low frequency maxima. The average spectra in the recirculation
region are shown in Figs. 12c and 12d for = 20 deg, Rec
= 1.47 104 and 2.67 104. A broad low frequency maximum occurs near Sth = 0.2 for both values of Rec, which is similar to the
051202-8 / Vol. 131, MAY 2009
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Fig. 7 Detected vortex positions shown for the first 200 ms of a total of
2000 ms with and without filtering. Unfiltered detected vortex positions for
= 0.70 at =14 deg and a Rec = 1.47 104, b Rec = 2.67 104, and c
Rec = 6.67 104. POD reconstructed detected vortex positions filtered for
path lifetime greater than 0.05c for d Rec = 1.47 104, e Rec = 2.67 104,
and f Rec = 6.67 104 at = 14 deg.
velocity and frequency data. The two length scales are shown to
be within 10% for all values for Rec = 1.47 104 and 4% for
Rec = 2.67 104, with no observable trend with angle of attack for
the range studied. Based on the variability of U and the resolution of the peak frequencies, the results for both length scales
show very good agreement. The length scale is shown to decrease
with increasing Rec, which indicates a relative increase in rate of
vortex generation within the shear layer as Rec increases.
/ U
U / U
deg
Rec
14,700
26,700
66,700
14,700
26,700
66,700
4
16
18
0
0.44
0.41
0.42
0.45
0.36
0.38
0.35
0.33
0.23
0.22
0.21
0.20
0.41
0.40
0.37
0.42
0.40
0.42
0.41
0.38
0.46
0.49
0.49
0.50
Avg.
0.43
0.36
0.22
0.40
0.40
0.49
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Fig. 9 a rms value of , , for the Rec = 1.47 104, = 20 deg case.
The solid white line identifies the region of highest amplitude. b rms value
of the temporal autocorrelation of , . The two areas outlined with
dashed lines represent regions of large amplitude cyclic activity of .
Conclusions
Fig. 10 Average spectral results using four points in the high cyclic shear
region shown in Fig. 9b for Rec = 1.47 104, = 20 deg. TR PIV data at 500
Hz over 2 s: a averaged spectrum of v, b averaged spectrum of u, c
averaged spectrum of , and d hot wire data spectrum for a single point in
the shear region using data at 10 kHz for 10 s.
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Hz. Lower modes of the POD of the velocity field were used to
reconstruct a low-pass filtered velocity field, which assisted in
tracking vortices. A vortex convection velocity was determined
for three relatively low Reynolds numbers and was not seen to
Table 2 Length scales based on i the average separation distance between tracked vortices, ltrack, and ii the frequency and
convection velocity of detected vortices, lfreq
Rec
deg
ltrack / c
lfreq / c
ltrack / c
lfreq / c
14
16
18
20
0.034
0.036
0.038
0.032
0.040
0.038
0.035
0.041
0.025
0.028
0.032
0.028
0.030
0.028
0.028
0.023
Avg.
0.035
0.038
0.028
0.027
14,700
26,700
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Acknowledgment
Funding for this work was partially provided by AFOSR under
Grant No. FA-9550-05-1-0041 and is gratefully acknowledged.
Nomenclature
A
AM
c
D
h
lfreq
ltrack
M
N
PM
Rec
S
St
Sth
U
UM
U
U *
w
V
Z
Greek Notation
M
tc
l
References
1 McCullough, G. B., and Gault, D. E., 1951, Examples of Three Representative Types of Wing-Section Stall at Low Speeds, NACA Technical Note 2502.
2 Carmichael, B. H., 1981, Low Reynolds Number Wing Survey, Volume 1,
NASA Contractor Report No. CR-165803.
3 Re, R. J., Pendergraft, O. C., Jr., and Campbell, R. L., 2006, Low Reynolds
Number Aerodynamic Characteristics of Several Airplane Configurations Designed to Fly in the Mars Atmosphere at Subsonic Speeds, NASA Report No.
214312.
4 Mueller, T. J., and DeLaurier, J. D., 2003, Aerodynamics of Small Vehicles,
Annu. Rev. Fluid Mech., 35, pp. 89111.
5 Torres, G. E., and Mueller, T. J., 2001, Aerodynamic Characteristics of Low
Aspect Ratio Wings at Low Reynolds Numbers, Prog. Astronaut. Aeronaut.,
195, pp. 115142.
6 Broeren, A. P., and Bragg, M. B., 2001, Unsteady Stalling Characteristics of
Thin Wings at Low Reynolds Number, Prog. Astronaut. Aeronaut., 195, pp.
191213.
7 Bishop, R. E. D., and Hassan, A. Y., 1964, The Lift and Drag Forces on a
Circular Cylinder in a Flowing Fluid, Proc. R. Soc. London, Ser. A, 277, pp.
3250.
8 Berger, E., and Wille, R., 1972, Periodic Flow Phenomena, Annu. Rev. Fluid
Mech., 4, pp. 31340.
9 Mair, W. A., and Maull, D. J., 1971, Bluff Bodies and Vortex SheddingA
Report on Euromech 17, J. Fluid Mech., 452, pp. 209224.
10 Yarusevych, S., Sullivan, P. E., and Kawall, J. G., 2005, Wing Boundary
Layer Separation and Control at Low Reynolds Numbers, Exp. Fluids, 38,
pp. 545547.
11 Tinar, E., and Cetiner, O., 2006, Acceleration Data Correlated With PIV
Images for Self-Induced Vibrations of a Wing, Exp. Fluids, 41, pp. 201212.
12 Cheng, M., and Chen, B. K., 2007, A Numerical Study on Fluid Force Reduction of a Square Cylinder by Flow Control, Fluids Engineering Conference, San Diego, CA, Jul. 30Aug. 2, Paper No. FEDSM2007-37025.
13 Roshko, A, 1954, On the Drag and Shedding Frequency of Two Dimensional
Bluff Bodies, NACA Technical Note 3169.
14 Abernathy, F. H., 1962, Flow Over an Inclined Plate, ASME J. Basic Eng.,
84, pp. 380388.
15 Burgmann, S., Brcker, C., and Schrder, W., 2006, Scanning PIV Measurements of a Laminar Separation Bubble, Exp. Fluids, 41, pp. 319326.
16 Derksen, R. W., and Rimmer, J., 2006, Aerodynamic Flow Simulation, Int.
Conference on Advances in Fluid Mechanics VI, Skiathos, Greece, pp. 5967.
17 Jeong, J., and Hussain, F., 1995, On the Identification of a Vortex, J. Fluid
Mech., 285, pp. 6994.
18 Adrian, R. J., Christiansen, K. T., and Liu, Z.-C., 2000, Analysis and Interpretation of Instantaneous Velocity Fields, Exp. Fluids, 29, pp. 275290.
19 Graftieaux, L., Michard, M., and Grosjean, N., 2001, Combining PIV, POD
and Vortex Identification Algorithms for the Study of Unsteady Turbulent
Swirling Flows, Meas. Sci. Technol., 12, pp. 14221429.
20 Shinneeb, A.-M., Balachandar, R., and Bugg, J. D., 2008, Analysis of Coherent Structures in the Far-Field Region of an Axisymmetric Free Jet Identified
Using Particle Image Velocimetry and Proper Orthogonal Decomposition,
ASME J. Fluids Eng., 130, p. 011202.
21 Agrawal, A., and Prasad, A. K., 2002, Properties of Vortices in the SelfSimilar Turbulent Jet, Exp. Fluids, 334, pp. 565577.
22 Troolin, D. R., Longmire, E. K., and Lai, W. T., 2006, Time Resolved PIV
Analysis of Flow Over a NACA 0015 Airfoil With Gurney Flap, Exp. Fluids,
41, pp. 241254.
23 Druault, P., Guibert, P., and Alizon, F., 2005, Use of Proper Orthogonal Decomposition for Time Interpolation From PIV Data, Exp. Fluids, 39, pp.
10091023.
24 Weiland, C., and Vlachos, P., 2007, Analysis of the Parallel Blade Vortex
Interaction With Leading Edge Blowing Flow Control Using the Proper Orthogonal Decomposition, ASME Fluids Engineering Summer Conference,
San Diego, CA, Paper No. FEDSM2007-37275.
25 Kurtulus, D. F., Scarano, F., and David, L., 2007, Unsteady Aerodynamic
Forces Estimation on a Square Cylinder by TR-PIV, Exp. Fluids, 42, pp.
185196.
26 Hinze, J. O., 1959, Turbulence, McGraw-Hill, New York.
27 Dano, B. P. E., and Liburdy, J. A., 2006, Vortical Structures of a 45 Inclined
Pulsed Jet in Crossflow, Fluid Dynamics Conference, San Francisco, CA,
AIAA Paper No. 2006-3543.
28 Morse, D. R., and Liburdy, J. A., 2007, Dynamic Characteristics of Flow
Separation from a Low Reynolds Number Wing, Fluids Engineering Conference, San Diego, CA, Jul. 30Aug. 2, Paper No. FEDSM20070-37083.
29 Chen, G., Lin, Z., Morse, D., Snider, S., Apte, S., Liburdy, J., and Zhang, E.,
2008, Multiscale Feature Detection in Unsteady Separated Flows, IJNAM,
5, pp. 1735.
30 Chatterjee, A., 2000, An Introduction to the Proper Orthogonal Decomposition, Curr. Sci., 787, pp. 808817.
31 Sirovich, L., 1987, Turbulence and the Dynamics of Coherent Structures
Part I: Coherent Structures, Q. Appl. Math., 453, pp. 561571.
32 Berkooz, G., Holmes, P., and Lumley, J. L., 1993, The Proper Orthogonal
Decomposition in the Analysis of Turbulent Flows, Annu. Rev. Fluid Mech.,
25, pp. 539575.
33 Morse, D. R., and Liburdy, J. A., 2008, Experimental Time Resolved Flow
Features of Separation Over an Elliptic Leading Edge, 46th AIAA Aerospace
Sciences Meeting and Exhibit, Reno, NV, Jan. 710, AIAA Paper No. 2008655.
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
The mean velocity profile and friction factor in turbulent flows with polymer additives are
investigated using Prandtls mixing-length theorem. This study reveals that the mixinglength theorem is valid to express the drag-reducing phenomenon and that the presence
of polymer additives increases the damping factor B in van Driests model; subsequently
reducing the mixing-length, this interprets that the polymer hampers the transfer of turbulent momentum flux, the velocity is increased, and flow drag is reduced. This study also
discusses the onset Reynolds number for drag reduction to occur. The predicted velocity,
friction factor, and onset Reynolds number are in good agreement with the measured data
in the literature. DOI: 10.1115/1.3111255
Keywords: drag reduction, velocity distribution, van Driests model, damping factor,
friction factor
Introduction
du
+ uv
dy
du
dy
2a
eff = uh
2b
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
du
du
y
= uv + + pRyy = u2 1
dy
dy
h
where Ryy is the ensemble average dimensionless extension tensor of the dyadic product of the end-to-end vector distance of the
polymer chains in the y direction and p is a viscosity parameter
that is related to the concentration of the polymer. The right-hand
side of Eq. 3 is simply the rate at which momentum is produced
by the pressure head, and on the left-hand side the Reynolds stress
is the momentum flux 2.
Gyr and Tsinobers 19 expression shown in Eq. 1 can be
rewritten as follows:
du
y
du
= uv + + eff = u2 1
dy
dy
h
Equation 5 establishes the relationship between the effective viscosity and properties of polymer solution in drag-reducing flows.
If Ryy and p can be determined from the polymer concentration
and the vector distance of the polymer molecules, then one is able
to determine eff. Equation 4 provides a useful tool to assess the
effective viscosity eff from the measured Reynolds shear stress
and velocity gradient. In other words, if the FENE-P model is
correctly expressed as the interaction of polymer molecules and
turbulence, then the obtained pRyy from the property of polymer and concentration must be the same as eff obtained from the
properties of turbulence using Eq. 4; this conclusion can be extended to other models, such as Oldroyd-B model and Maxwell
model.
Instead of solving Eq. 3 directly, Benzi et al. 2 and Lvov et
al. 16 introduced the energy balance equation and other assumptions, such as u21 y / h u2 and uv 0, then they claimed
that the velocity equation they obtained is valid even for the
whole flow region.
Different from the approach used by Benzi et al. 2 and Lvov
et al. 16, the present study makes an attempt on the direct solution of Eq. 3. By inserting Eq. 2b into Eq. 4, one obtains the
dimensionless Reynolds shear stress
u v
du+
y+
=
1
h+
dy +
u2
where u+ = u / u, y + = uy / , h+ = uh / , and
D = 1 + h +
eff = pRyy
Yang and Dou 22 investigated the velocity profiles in turbulent boundary layer flows; the emphasis of this study will be
placed on the pipe flows. The objectives of this study are as follows: 1 to theoretically investigate the mean velocity profile and
friction factor of drag reduction flow, 2 to quantitatively express
the onset Reynolds number for drag reduction, and 3 to determine the relation between the damping factor in van Driests 27
model and D.
u v = l 2
du
dy
where l is the mixing-length over which the turbulence characteristics remain unchanged 28. Inserting Eq. 9 into Eq. 6, one
has
l+2
du+
dy +
+ D
du+
1 = 0
dy +
10
11
u =
y+
21
dy
D + D2 + 4l+21
12
13
where k is the von Karman constant. y is van Driests damping function, which can be expressed as follows:
+
y + = 1 exp y +/B
14
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
u h
0.1
15
16
It can be seen clearly from Fig. 1 that Virks three-layer model can
be expressed by a single equation, i.e., Eq. 12.
In Fig. 1, the dotted line represents Virks asymptote of maximum drag reduction MDR, i.e.,
Journal of Fluids Engineering
u
uy
= 11.7 ln
17.0
u
17
Bo
18
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
lar weight and intrinsic viscosities in their experiments, the parameter D is estimated by the best-fit to experimental data instead
of using Eq. 8, and the damping factor B is calculated from Eq.
18 based on D.
It can be seen from Fig. 3 that the agreements between the
measured velocity profiles in drag-reduced flows and Eq. 12 are
reasonable. Virks asymptote is also included in Fig. 3 for comparison. From Fig. 3, one finds that Virks asymptote does not
represent a universal relationship for the maximum drag reduction; the experimental data of Ptasinski et al. 25,35 begin to
exceed Virks asymptote significantly at y + = 80. Warholic et al.
18 also concluded that Virks asymptote may not be universal.
However, it can be seen from Figs. 1 and 3 that Eq. 12 can
represent the data points. In the turbulent core, the solid lines
shown in Figs. 1 and 3 are parallel to the log-law for Newtonian
fluid flows with a shift of B without change in slope, as stated by
Eq. 16.
This study assumes that the velocity distribution in the viscous
sublayer follows Eq. 11 because in the near wall region the
distribution of the velocity must follow u+ = y +, not u+ = y + / D as
Eq. 11 gives. Thus it is necessary to test the hypothesis using the
velocity data in the near wall region. Figure 4 shows the measured
velocity distribution inside the viscous sublayer. It can be seen
clearly from Fig. 4 that this model can express the velocity distribution very well, indicating that the assumption is reasonable.
Currently, many researchers used direct numerical simulation
DNS to model the mechanism of drag-reducing flow, such as De
Angelis et al. 36, Min et al. 4,5, Housiadas and Beris 37, and
Ptasinski et al. 25, and their results show that DNS is a powerful
tool to understand the drag-reducing flow. This study uses a different approach to express the drag-reducing effect; thus it would
be useful to compare the proposed model with DNS method.
Source
Den Toonder et al. 17
Luchik and Tierderman 12
Warholic et al. 8
Harder and Tiederman 13
Warholic et al. 18
Ptasinski et al. 25
Solvent
Polymer
Water
Water
Water
Water
Water
Water
Superfloc A110
Separan AP-273
Percol 727
Separan AP-273
Percol 727
Superfloc A110
Polymer
Pipe radius/water
concentration,
depth,
Drag reduction
C wppm
h cm
DR
20
1.32.1
1.24
5
13
175, 435
4
1.25
2.54
3.0
2.54
4.0
24.2%
22.0%
42.0%
64.0%
65%, 70%
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Ptasinski et al. 35 measured the velocity profiles in dragreducing flows, and the comparison of the measured data with the
present model is included in Fig. 3, in which the discrepancy is
noticeable. Ptasinski et al. 25 modeled their measurements in
2001 using DNS, and the comparison between the same measured
data as shown in Fig. 3 with DNSs results is included in Fig. 5.
Based on the data of Ptasinski et al. 35 in Figs. 3 and 5, one
could find that the present model yields the best agreement relative to DNSs results as well as Virks asymptote of maximum
drag reduction.
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Reynolds number. In other words, the straight line in Fig. 7 interprets that polymer species and concentration only affect the value
of . For Magnifloc 837A at 100 ppm used by Reischman and
Tiederman 38, the coefficient = 0.0011. Therefore, one can
conclude from Reischman and Tiedermans 38 measurements
that the effective viscosity is proportional to uh, as shown in Eq.
2b.
Friction Factor
Equations 7 and 8 state that the elastic factor is independent of flow strength, and D varies linearly with the Reynolds
number, i.e., ur / ; this has been proved in Fig. 7. It would be
interesting to investigate how the factor changes with the polymer concentration. To do so, the friction factor of drag-reducing
flows f defined as follows is investigated:
u
f=2
V
u2 1
dul
y
=
dy
h
20
ul+ = y + 1
21
For the transitional region from the laminar to the turbulent state,
Yang and Dou 22 obtained the following velocity distribution:
ut+ = rlul+ + 1 rlu+
22
19
Fig. 7 Linear relation between D and Reynolds number based on Reischman and Tiedermans 38 data
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
1
rl = e
n=1
n Rk
n! R
2n
for R Rk
for R Rk
23
where R is the Reynolds number =uh / and Rk is the threshold Reynolds number at which the flow becomes unstable and
turbulence appears. For a pipe flow Rk = 67.82 or 2Vh / = 2300.
Therefore, one is able to obtain the drag coefficient f from the
mean velocity, which is determined by integrating Eq. 12,
1 1 V
1
R 1 rl
=
rl +
=
4
f 2 u 2
R2
2R y +u+dy +
24
The comparison of 1 / f given by Eq. 24 and Virks 39 experimental data in smooth pipes is presented in Fig. 8, which shows
that Eq. 24 = 0 represents the data points obtained in Newtonian fluid flows well. For drag-reducing flows, Virk used polyethyleneoxide N750 o = 0.019 as the polymer additive with the
following polymer concentrations: 43.6 ppm, 98.6 ppm, and 939
ppm; Eq. 8 gives the corresponding elastic factors
= 0.000158, 0.000326, 0.000817. In Fig. 8, the experimental data
for polyethyleneoxide W301 o = 0.166 with a concentration of
18.7 ppm are also included for comparison; Eq. 8 gives
= 0.0033 for this case. Figure 8 indicates that for each line, the
parameter holds constant over the full range of Reynolds numbers, i.e., uh / . The good agreement between the measured and
predicted friction factors indicates that is independent of the
Reynolds number.
For the drag reduction flows shown in Fig. 8, the calculated
onset Reynolds number Re f=22uh / from Eq. 15 is 1790
for = 0.000158, which is very close to the measured value of
1500; for = 0.000326, the calculated Re f is 867 and the observed value is 960; thus, Eq. 15 can express the onset of drag
reduction 4043.
Conclusions
The applicability of Prandtls mixing-length theorem in dragreducing flows is investigated, and the mixing-length equation
developed by van Driest is used. This study finds that the damping
factor B in van Driests model depends on the effective viscosity
caused by polymer additives, and the velocity in visco-elastic fluid
flow can be well predicted by Prandtls theorem. Based on the
comparison between measured and predicted velocity and friction
factors in pipe/channel flows, this study reaches the following
conclusions.
Journal of Fluids Engineering
1. The difference between the drag-reducing flow and Newtonian fluid flow is caused by the stress deficit, which can be
well represented by the sole parameters, D or . The experimental data confirm that D is proportional to the Reynolds number.
2. The mixing-length model can be applied to drag-reducing
flows, and the mixing-length is reduced after the polymer is
added to the Newtonian fluid flow, or the damping factor in
van Driests expression increases with D. The calculated
results show that the mean velocity becomes higher with the
increase in D. The measured velocity profiles can be described by Eq. 12 from the viscous sublayer through the
buffer layer to the turbulent core.
3. The measured friction factor shows that the elastic factor
or effective viscosity eff is independent of Reynolds number
ur / . The friction factor can be expressed by the derived
equation.
4. The theoretical results show that the onset Reynolds number
only depends on the drag-reduction parameter . The onset
Reynolds number decreases with the increase in . If D
1.1, the velocity profile and friction factor for dragreducing flows perform as the Newtonian fluid flow.
Acknowledgment
The author would like to express his sincere appreciation to the
anonymous referees for their careful review that greatly improves
this papers quality.
References
1 Toms, B. A., 1949, Some Observation on the Flow of Linear Polymer Solutions Through Straight Tubes at Large Reynolds Numbers, Proceedings of the
First International Congress on Rheology, Vol. 2, pp. 135141.
2 Virk, P. S., 1971, An Elastic Sublayer Model for Drag Reduction by Dilute
Solutions of Linear Macromolecules, J. Fluid Mech., 45, pp. 417440.
3 Larson, R. G., 2003, Analysis of Polymer Turbulent Drag Reduction in Flow
Past a Flat Plate, J. Non-Newtonian Fluid Mech., 11123, pp. 229250.
4 Min, T., Choi, H., and Yoo, J. Y., 2003, Maximum Drag-Reduction in a
Turbulent Channel Flow by Polymer Additives, J. Fluid Mech., 492, pp.
91100.
5 Min, T., Yoo, J. Y., Choi, H., and Joseph, D. D., 2003, Drag Reduction by
Polymer Additives in a Turbulent Channel Flow, J. Fluid Mech., 486, pp.
213238.
6 Gasljevic, K. Aguilar, G., and Matthys, E. F., 2001, On Two Distinct Types of
Drag-Reducing Fluids, Diameter Scaling, and Turbulent Profiles, J. NonNewtonian Fluid Mech., 963, pp. 405425.
7 Ryskin, G., 1987, Turbulent Drag Reduction by Polymers: A Quantitative
Theory, Phys. Rev. Lett., 59, pp. 20592062.
8 Warholic, M. D., Heist, D. K., Katcher, M., and Hanratty, T. J., 2001, A Study
With Particle-Image Velocimetry of the Influence of Drag-Reducing Polymers
on the Structure of Turbulence, Exp. Fluids, 31, pp. 474483.
9 White, C. M., Somandepalli, V. S. R., and Mungal, M. G., 2004, The Turbu-
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
lence Structure of Drag-Reduced Boundary Layer Flow, Exp. Fluids, 36, pp.
6269.
Durst, F., Keck, T., and Kline, R., 1985, Turbulence Quantities and Reynolds
Stresses in Pipe Flow of Polymer Solutions, Proceedings of the First International Conference on Laser Anemometer-Advances and Applications, BHRA,
p. 31.
Willmarth, W. W., Wei, T., and Lee, C. O., 1987, Laser Anemometer Measurements of Reynolds Stress in a Turbulent Channel Flow With Drag Reducing Polymer Additives, Phys. Fluids, 30, pp. 933.
Luchik, T. S., and Tierderman, W. G., 1988, Turbulent Structure in LowConcentration Drag-Reducing Channel Flows, J. Fluid Mech., 190, pp. 241
263.
Harder, K. J., and Tiederman, W. G., 1991, Drag Reduction and Turbulent
Structure in Two-Dimensional Channel Flows, Philos. Trans. R. Soc. London,
Ser. A, 336, pp. 1934.
Wei, T., and Willmarth, W. W., 1992, Modifying Turbulent Structure With
Drag-Reducing Polymer Additives in Turbulent Channel Flows, J. Fluid
Mech., 245, pp. 619641.
Lumley, J. L., 1969, Drag Reduction by Additives, Annu. Rev. Fluid Mech.,
1, pp. 367384.
Lvov, V. S., Pomyalov, A., Procaccia, I., and Tiberkvich, V., 2004, Drag
Reduction by Polymers in Wall Bounded Turbulence, Phys. Rev. Lett., 92, p.
244503-4.
Den Toonder, J. M. J., Hulsen, M. A., Kuiken, G. D. C., and Nieuwstadt, F. T.
M., 1997, Drag Reduction by Polymer Additives in a Turbulent Pipe Flow:
Numerical and Laboratory Experiments, J. Fluid Mech., 337, pp. 193231.
Warholic, M. D., Massah, H., and Hanratty, T. J., 1999, Influence of DragReducing Polymers on Turbulence: Effects of Reynolds Number, Concentration and Mixing, Exp. Fluids, 27, pp. 461472.
Gyr, A., and Tsinober, A., 1997, On the Rheological Nature of Drag Reduction Phenomena, J. Non-Newtonian Fluid Mech., 73, pp. 153162.
Schmmer, P., and Thielen, W., 1980, Structure of Turbulence in Viscoelastic
Fluids, Chem. Eng. Commun., 4, pp. 593606.
Giesekus, H., 1981, Structure of Turbulence in Drag Reducing Fluids, Lecture
Series 1981-86, von Karman Institute for Fluid Dynamics, Rhode-SainGense, Belgium.
Yang, S. Q., and Dou, G., 2005, Drag Reduction in Flat-Plate Turbulent
Boundary Layer Flow by Polymer Additive, Phys. Fluids, 176, p. 065104.
Dou, G. R., 1996, Basic Law in Mechanics of Turbulent Flows, China Ocean
Eng., 101, pp. 144.
Berman, N. S., 1989, Drag Reduction in Fluids Flows: Techniques for Friction
Control, R. H. J. Sellin and R. T. Moses, eds., Ellis Horwood Limited, Chichester.
Ptasinski, P. K., Boersma, B. J., Nieuwstadt, F. T. M., Hulsen, M. A., Van den
Brule, B. H. A. A., and Hunt, J. C. R., 2003, Turbulent Channel Flow Near
Maximum Drag Reduction: Simulations, Experiments and Mechanisms, J.
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Shuhong Liu1
e-mail: liushuhong@mail.tsinghua.edu.cn
Liang Zhang
State Key Laboratory of Hydro Science and
Hydraulic Engineering,
Tsinghua University,
Beijing 100084, China
Michihiro Nishi
Department of Mechanical Engineering,
Kyushu Institute of Technology,
Sensui-cho 1-1,
Tobata, Kitakyushu 804-8550, Japan
e-mail: nishi@mech.kyutech.ac.jp
Yulin Wu
State Key Laboratory of Hydro Science and
Hydraulic Engineering,
Tsinghua University,
Beijing 100084, China
Introduction
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
u j = 0
+
t xj
caca
cacau j = S
+
t
xj
ui
p ji
u j u i = g i
+
+
t
xj
xi xi
2
r
3ca
M
C1
r
2R
1/2
pv
T
31 ca u
M
C2
r
2R
1/2
pv
T
= caca + uu + 1 ca ul
where ca = v is the density of saturated vapor and u is the density of nondissolved gases, both of which are in the following
forms from ideal gas assumption:
ca =
M p
R T
u =
M u p
R T
10
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
= caca + uu + 1 ca ul
11
where ca is the viscosity of cavity phase, l is the water viscosity, and u is the viscosity of nondissolved gas in water.
2.4 Boundary Conditions for Steady Flow. The following
boundary conditions are used to get steady solution of cavity/
liquid turbulent flow in the present simulation.
1 Input data of all physical variants are prescribed at the inlet
boundaries of the flow field.
2 The gradients of all physical variants normal to the boundaries are given at the outlet.
3 On solid walls of the domain, the nonslip flow condition is
adopted. The velocity distribution in the boundary layer is
expressed as the wall function near the solid walls.
4 Zero pressure gradient normal to the surface is used for all
boundaries of the domain except a point corresponding to
the reference pressure. In order to satisfy the continuity
condition from the inlet to the outlet of calculation domain,
correction of velocity magnitude to the outlet distribution is
made based on the difference between the inflow and the
outflow during the calculation.
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Value
ZS = 24
Z0 = 24
D0 = 1.111D1
B0 = 0.288D1
D1 = 372.2 mm
D2 = 366.0 mm
Z = 15
Fig. 5 Model Francis turbine: a calculation domain, b interaction surfaces, and c survey points of pressure in draft tube
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Total
elements
Runner
elements
Torque
N m
1
2
3
1,180,000
1,880,000
3,710,000
100,000
780,000
2,630,000
6
3
2
1129.792
1149.769
1149.767
Cases
Ha HVA HS HV
H
12
where HVA is the vacuum head in the draft tank of the test rig,
which indicates the pressure level at the outlet of the draft tube.
Ha is the atmospheric pressure head. HS is the suction head of the
turbine. H is the test head acting on the turbine. HV is the vapor
pressure head at the test temperature.
In the cavitating flow simulation, the following assumptions
were used:
1 radius of cavitation nuclei, r0 = 0.5 m
2 volume fraction of nondissolved gas, u = 5 105
3 temperature at the inflow plane, T = 298 K
4.2 Validation of Simulation. Since applicability of the cavitation model to the performance prediction of a hydroturbine is
our major concern, its validation is conducted from the steady
flow simulation.
Predicted torque. Table 4 shows the predicted values of runner
torque with and without cavities for three cases. We see that the
differences between the results of the single-phase simulation and
those of the cavitating flow are almost negligible. This is because
of the higher cavitation number around = 0.142 at the corresponding operating condition. Though the underestimation is ob-
Experiment
N m
Single-phase
N m
Cavitating flow
N m
744.5
1178.4
1450.3
682.3
1149.7
1393.6
683.9
1149.7
1395.3
Cases
Small flow rate
= 0.143
Optimum
= 0.144
Large flow rate
= 0.132
GV opening
mm
Head
m
Flow rate
m3 / s
Rotational speed
min1
Q11
m3 / s
n11
rpm
10
30
0.322
1122
0.466
75
15
30
0.497
1122
0.678
75
22
30
0.647
1122
0.882
75
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
of
efficiency
Q11
such operating points. To see the usability of the present numerical method for predicting the pressure fluctuation, Case 1 in Table
3 is treated as a typical example under the cavitation condition of
= 0.07. Here, the time step of 0.001 s was adopted in the unsteady turbulent flow analysis. Thus, the time step corresponds to
the runner rotation of 6.72 deg, as the rotational speed of the
runner is 1122 min1. Following IEC-60193, peak-to-peak amplitude and fundamental frequency of pressure fluctuation are evaluated at two survey points, as shown in Fig. 5c. One of them is
expressed as the downstream point 0.3D1, and the other is the
downstream point 1.0D1. They are located at the distances of 0.3
and 1.0 times the runner inlet diameter from the runner exit,
respectively.
The predicted results of dimensionless peak-to-peak amplitude
Conclusions
1 The critical cavitation number c and hydraulic performance of a Francis turbine are reasonably predicted by the
steady cavitating flow simulation based on a mixture
model.
2 The cavitated vortex rope in the draft tube at overload operation was reasonably displayed from the numerical re-
Single-flow calculation
Survey point
Amp. H / H
%
Fundamental
freq. Hz
Amp. H / H
%
Fundamental
freq. Hz
Amp. H / H
%
Fundamental
freq. Hz
0.3D1
1.0D1
3.91
3.66
3.43
3.42
4.32
4.1
4.98
4.98
4.25
3.2
4.12
4.12
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Acknowledgment
The research work was funded by Chinese National Foundation
of Natural Science Contract No. 90410019.
Nomenclature
i
c
ji
correction coefficients
gravitational acceleration
test head acting on the turbine
atmospheric pressure head
suction head of the turbine
saturated vapor pressure head
vacuum head in the draft tank
kinetic energy of turbulence
Moore molecule weight
rotational speed
unit speed, =nD / H
local pressure
flow rate
unit flow rate, =Q / D2H
gas constant
radius of the bubble
mass transfer rate
temperature
time
velocity
coordinates
volume fraction
surface tension
turbulent kinetic energy dispassion rate
efficiency
viscosity
density
cavitation number, see Eq. 12
initial cavitation number
critical cavitation number
stress
Subscripts
ca
l
u
v
cavity phase
liquid
undissolved gas
vapor
C 1, C 2
g
H
Ha
HS
HV
HVA
k
M
n
n11
p
Q
Q11
R
r
S
T
t
u
x,y,z
Performance
Unit
Measuring range
Uncertainty
1
2
3
4
Flow rate Q
Head H
Force on arm K
Angular speed
m3 / s
MPa
E pound
rad/s
01.0
02.07
01000
03000
0.20%
0.075%
0.02%
0.05%
References
1 Susan-Resiga, R. F., Muntean, S., and Anton, I., 2002, Numerical Analysis of
Cavitation Inception in Francis Turbine, Proceedings of the 21st IAHR Symposium on Hydraulic Machinery and Systems, F. Avellan, G. Ciocan, and S.
Kvicinsky, eds., Lausanne, Switzerland, pp. 431438.
2 Yamaguchi, H., and Kato, H., 1983, On Application of Nonlinear Cavity
Flow Theory to Thick Foil Section, Proceedings of the Second International
Conference of Cavitation, ImechE, Paper No. C209/83, pp. 167174.
3 Brewer, W. H., and Kinnas, S. A., 1995, Experimental and Computational
Investigation of Sheet Cavitation on a Hydrofoil, Proceedings of the Second
Joint ASME/JSME Fluids Engineering Conference and ASME/EALA Sixth
International Conference on Laser Anemometry, Hilton Head Island, SC, pp.
115.
4 Pellone, C., and Peallat, J. M., 1995, Non-Linear Analysis of ThreeDimensional Partially Cavitating Hydrofoil, Proceedings of the International
Symposium on Cavitation, Deauville, France, pp. 6367.
5 De Lange, D. F., and De Bruin, G. J., 1998, Sheet Cavitation and Cloud
Cavitation Re-Entrant Jet and Three-Dimensionality, Appl. Sci. Res., 581
4, pp. 91114.
6 Watanabe, S., Hidaka, T., Horiguchi, H., Furukawa, A., and Tsujimoto, Y.,
2007, Steady Analysis of the Thermodynamic Effect of Partial Cavitation
Using the Singularity Method, ASME J. Fluids Eng., 1292, pp. 121127.
7 Chen, Y., and Heister, S. D., 1994, A Numerical Treatment for Attached
Cavitation, ASME J. Fluids Eng., 116, pp. 613618.
8 Ventikos, Y., and Tzabiras, G., 2000, A Numerical Method for the Simulation
of Steady and Unsteady Cavitating Flows, Comput. Fluids, 29, pp. 6388.
9 Horiguchi, H., Arai, S., Fukutomi, J., Nakase, Y., and Tsujimoto, Y., 2004,
Quasi-Three-Dimensional Analysis of Cavitation in an Inducer, ASME J.
Fluids Eng., 126, pp. 709715.
10 Arndt, R. E. A., Kjeldsen, M., Song, C. C. S., and Keller, A., 2002, Analysis
of Cavitation Wake Flows, Proceedings of the 21st IAHR Symposium on
Hydraulic Machinery and Systems, F. Avellan, G. Ciocan, and S. Kvicinsky,
eds., Lausanne, Switzerland, pp. 395402.
11 Qin, Q., Song, C. S. S., and Arndt, R. E. A., 2003, Numerical Study of
Unsteady Turbulent Wake Behind a Cavitating Hydrofoil, Fifth International
Symposium on Cavitation, Osaka, Japan, Paper No. EM 003.
12 Deshpande, M., Feng, J., and Merkle, C. L., 1997, Numerical Modeling of the
Thermodynamic Effects of Cavitation, ASME J. Fluids Eng., 119, pp. 420
427.
13 Rieger, R., 1992, Mehrdimensionale Berechnung Zweiphasiger Stroemungen, Ph.D. thesis, Technical University Graz, Graz, Austria.
14 Grogger, H. A., and Alajbegovic, A., 1998, Calculation of the Cavitating
Flow in Venturi Geometries Using Two Fluid Model, Proceedings of the
ASME Fluids Engineering Division Summer Meeting, Washington, DC, Paper
No. FEDSM98-5295.
15 Liu, S. H., Wu, Y. L., and Luo, X. W., 2005, Numerical Simulation of 3D
Cavitating Turbulent Flow in Francis Turbine, Proceedings of the ASME
FEDSM, Houston, TX, Paper No. FEDSM2005-77017.
16 Singhal, A. K., Vaidya, N., and Leonard, A. D., 1997, Multi-Dimensional
Simulation of Cavitating Flows Using a PDF Model for Phase Change, Proceedings of the ASME Fluids Engineering Division Summer Meeting, Vancouver, BC, Vol. 4, pp. 18.
17 Kunz, R. F., Boger, D. A., Stinebring, D. R., Chyczewski, T. S., Lindau, J. W.,
Gibeling, H. J., Venkateswaran, S., and Govindan, T. R., 2000, A Preconditioned Navier-Stokes Method for Two-Phase Flows With Application to Cavitation Prediction, Comput. Fluids, 29, pp. 849875.
18 Senocak, I., and Shyy, W., 2002, A Pressure-Based Method for Turbulent
Cavitating Flow Computation, J. Comput. Phys., 176, pp. 363383.
19 Singhal, A. K., Athavale, M. M., Li, H. Y., and Jiang, Y., 2002, Mathematical
Basis and Validation of the Full Cavitation Model, ASME J. Fluids Eng.,
1243, pp. 617624.
20 Aschenbrenner, T., Otto, A., and Moser, W., 2006, Classification of Vortex
and Cavitation Phenomena and Assessment of CFD Prediction Capabilities,
Proceedings. of the 23rd IAHR Symposium on Hydraulic Machinery and Systems, J. Kurokawa, ed., Yokohama, Japan, Paper No. F132.
21 Saito, Y., Takami, R., Nakamori, I., and Ikohagi, T., 2007, Numerical Analysis of Unsteady Behavior of Cloud Cavitation Around a NACA0015 Foil,
Comput. Mech., 40, pp. 8596.
22 Okita, K., and Kajishima, T., 2002, Numerical Simulation of Unsteady Cavitating Flow Around a Hydrofoil, Trans. Jpn. Soc. Mech. Eng., Ser. B, 68, pp.
637644.
23 Guo, Y., Kato, C., and Miyagawa, K., 2006, Large-Eddy Simulation of NonCavitating and Cavitating Flow in Venturi Geometries Using Two Fluid
Model, Proceedings of the 23rd IAHR Symposium on Hydraulic Machinery
and Systems, J. Kurokawa, ed., Yokohama, Japan, Paper No. F195.
24 Lindau, J. W., Kunz, R. F., Boger, D. A., Stinebring, D. R., and Gibeling, H. J.,
2002, High Reynolds Number, Unsteady, Multiphase CFD Modeling of Cavitating Flows, ASME J. Fluids Eng., 1243, pp. 607616.
25 Medvitz, R. B., Kunz, R. F., Boger, D. A., Lindau, J. W., and Yocum, A. M.,
2002, Performance Analysis of Cavitating Flow in Centrifugal Pumps Using
Multiphase CFD, ASME J. Fluids Eng., 1242, pp. 377383.
26 Fortes-Patella, R., Coutier-Delgosha, O., Perrin, J., and Reboud, J. L., 2007,
Numerical Model to Predict Unsteady Cavitating Flow Behavior in Inducer
Blade Cascades, ASME J. Fluids Eng., 1292, pp. 128135.
27 Xing, T., Li, Z. Y., and Frankel, S. H., 2005, Numerical Simulation of Vortex
Cavitation in a Three-Dimensional Submerged Transitional Jet, ASME J. Flu-
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
R. Asmatulu
Department of Mechanical Engineering,
Wichita State University,
1845 Fairmount,
Wichita, KS 67260-0133
S. Kim
F. Papadimitrakopoulos
Nanomaterials Optoelectronics Laboratory,
Polymer Program,
Institute of Materials Science,
University of Connecticut,
Storrs, CT 06269-3136
H. Marcus
Department of Materials Science and
Engineering,
Institute of Materials Science,
University of Connecticut,
Storrs, CT 06269-3136
Parallel-Plate Conductive
Electrodes for the Fabrication of
Larger 2D Colloidal Photonic
Crystals
A new dielectrophoretic force-induced parallel-plate assembly technique was used to
achieve close-packed 2D large colloidal photonic crystals on gold electrodes (200 nm
thick). The electrodes were patterned on a glass substrate using a conventional UV
lithography technique. The experimental tests conducted with 5.3 m carboxyl functionalized polystyrene particles at various ac and dc voltages, frequencies, and particle
concentrations showed that larger size 0.25 3 mm2 colloidal photonic crystals were
fabricated on the ground electrode rather than on the working electrode. To date, this is
the largest colloidal photonic crystal fabricated using this method. The reason behind this
phenomenon can be attributed to the electro-osmotic flow in the colloidal system and
dipole-dipole attractions between the colloidal particles. DOI: 10.1115/1.3111257
Keywords: parallel-plate gold electrodes, colloidal PS particles, 2D photonic crystals
Introduction
Theory
Dielectrophoretic force relies on the difference in the polarizability of the system compared with its surrounding fluid medium
e.g., water. Consider an isolated/dispersed spherical colloidal
particle or biological cell with known diameter and dielectric constant in a liquid medium. Under the nonuniform electric field, the
induced dipole moment drives the particle in a direction. The dielectrophoretic force acting on the spherical body is expressed as
13
FDEP = 2m ReKr3 Erms2
where r is the sphere radius, m is the permittivity of the suspending medium, Erms is the gradient of the root mean square electric
field, and ReK is the real part of the ClausiusMossotti factor
given by
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Fig. 1 Schematic outline of the experimental procedure left utilized to assemble larger 2D colloidal photonic crystals on the parallel-plate gold electrodes right connected to a power supplier. The red line is the working electrode, while the dark line is the ground electrode on which the photonic
crystals were formed not to be scaled.
K =
p m
2m
j
=
1 V202
8 a1 + 22
where V0 is the potential applied to the electrodes, is the viscosity of the electrolyte, and a is the distance from the center of
the electrode gap to the point of interest. In this model, the nondimensional frequency is given by
= r
Experimental
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
device was washed out several times using DI water. However, the
evidence indicated that some of the colloidal particles still remained on the ground electrodes. The reason for this may be
attributed to the excess amount of current/charge on the electrodes
and particles, which created a long-range attraction force to hold
the particles together. Figure 7 shows the crystalline PS particles
on the ground electrode photographed after washing with DI water.
4.4 Schematic Illustration of Crystallization Process. In the
electro-osmosis process high-velocity streamlines on the edge of
the electrodes, the charged particles tend to migrate toward the
lesser charged and turbulence areas as is specified in Eq. 5
2830. Thus, this phenomenon may possibly affect the colloidal
particles concentrated in specific locations. In the presence of
electrically charged ions in an aqueous system, the ions possibly
transfer onto the surface of the PS particles and increase the longrange dipole-dipole particle-particle interactions for the crystallization of the polarized particles 31. Based on the test results, it
is believed that this can be one of the major mechanisms of the
colloidal particle accumulation and crystallization on the ground
electrodes. Additional reasons may be electrohydrodynamic, electric field gradient, and thermal effects on the particles 2630.
Figure 8 depicts the possible schematic views of the larger PC
particle crystallization on the ground electrode using dc and ac
voltages.
Crystallization was not achieved on the working electrode,
which could be because of induced charges greatly increasing the
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Conclusions
The parallel-plate DEP tests were conducted on carboxyl functionalized PS colloidal particles 5.3 m to create larger 2D colloidal crystals. Test results revealed that colloidal particles could
be easily formed into closely packed larger colloidal crystals on
the ground electrodes for various dc and/or ac voltages. As is
known, the DEP technique is a simple, inexpensive, and more
rapid assembly technique compared with other methods. Therefore, this technique can have great potential for the photonic crystal fabrication, as well as many other biological, organic, and inorganic particle manipulations in the future.
Acknowledgment
The authors wish to acknowledge AFOSR Contract No.
F49620-01-1-0545 and ONR Contract No. N00014-00-1-0333
for their financial support.
References
1 Yablonovitch, E., 2001, Photonic Crystals: Semiconductors of Light, Sci.
Am., 2856, pp. 4755.
2 Johnson, S. G., and Joannopoulos, J. D., 2002, Photonic Crystals: The Road
From Theory to Practice, 1st ed., Kluwer, Dordrecht.
3 Park, S. H., and Xai, Y., 1999, Assembly of Mesoscale Particles Over Large
Areas and Its Application in Fabricating Tunable Optical Filters, Langmuir,
15, pp. 266273.
4 Kitaev, V., and Ozon, G. A., 2003, Self-Assembled Surface Patterns of Binary Colloidal Crystals, Adv. Mater. Weinheim, Ger., 15, pp. 7578.
5 Lumsdon, S. O., Kaler, E. W., and Volve, O. D., 2004, Two-Dimensional
Crystallization of Microspheres by a Coplanar AC Electric Field, Langmuir,
20, pp. 21082116.
6 Trau, M., Sankaran, S., Saville, D. A., and Aksua, I., 1995, Electric-FieldInduced Patterning of Colloidal Dispersions, Nature London, 374, pp. 437
439.
7 Mclachlan, M. A., Johnson, P. N., De La Rue, R., and McComb, D. W., 2004,
Thin Film Photonic Crystals: Synthesis and Characterization, J. Mater.
Chem., 14, pp. 144150.
8 Hamagami, J. H., Hasegawa, K., and Kanamura, K., 2004, Assembly of
Monodispersed Silica Spheres by Micro-Electrophoretic Deposition Process,
J. Ceram. Soc. Jpn., 12, pp. 169172.
9 Ahn, J. S., Hammond, P. T., Rubner, M. F., and Lee, I., 2005, Self-Assembled
Particle Monolayers on Polyelectrolyte Multiplayer: Particle Size Effects on
the Formation, Structure and Optical Properties, Colloids Surf., A, 259, pp.
4551.
10 Dziomkina, N. V., Hempenius, M. A., and Vancso, G. J., 2005, Symmetry
Control of Polymer Colloidal Monolayers and Crystals by Electrophoretic
Deposition Onto Patterned Surfaces, Adv. Mater. Weinheim, Ger., 172,
pp. 237240.
11 Mekis, A., Chen, J. C., Kurland, I., Fan, S., Villeneuve, P. R., and Joannopoulos, J. D., 1996, High Transmission Through Sharp Bends in Photonic Crys-
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Comparison of Turbulence
Modeling Strategies for Indoor
Flows
Ammar M. Abdilghanie
Lance R. Collins
David A. Caughey1
e-mail: dac5@cornell.edu
Sibley School of Mechanical & Aerospace
Engineering,
Cornell University,
Ithaca, NY 14853-7501
Turbulence modeling techniques are compared for the simulation of low speed indoor air
flow in a simple room. The effect of inlet turbulence intensity on the flow field is investigated using the constant coefficient large eddy simulation (LES) model with uniform
mean inlet conditions at several levels of inlet turbulence intensities. The results show
significant differences between the simulations with laminar inflow conditions and those
in which turbulence was introduced at the inlet. For simulations with turbulent inlet
conditions, it is noticed that the jet transitions to a state of fully developed turbulence
wherein the dynamics of the flow become nearly insensitive to any further increase in the
level of inlet turbulence. For laminar flow conditions, it is seen that the jet slowly spreads
and mixes with the quiescent room air. As a result, the jet reaches a fully developed
turbulent state further away from the inlet relative to the simulations with inlet turbulence. The effect of using experimental inlet profiles is also investigated. It is seen that,
close to the inlet, the flow is sensitive to the inflow details, whereas further away from the
inlet, these effects become less pronounced. The results from the constant coefficient and
the dynamic LES models are compared. The most noticeable differences in the flow occur
at the locations where the subgrid-scales contribution to the turbulent kinetic energy is
highest. Finally, the results from the dynamic LES and the k- models are compared. It is
found that there are significant differences between the two models for the zero inlet
turbulence limit where the flow is most probably transitional in nature and turbulence has
not yet reached a fully developed state. It is seen that in the laminar inflow case the k-
model predicts a fully turbulent jet very close to the inlet and thus fails to capture the
slow development of the jet found in LES. Accordingly, the k- model results are nearly
insensitive to the level of inlet turbulence especially far from the origin of the flow. It is
also seen that for cases with nonzero inlet turbulence level, the k- model predicts the
general features of the mean flow reasonably well; however, the k- model overpredicts
the jet spreading rate and the turbulent kinetic energy close to the inlet. Furthermore, the
k- model under predicts the turbulence level near the corner of the ceiling as it fails to
capture the complicated mean velocity and turbulent kinetic energy, most likely because
of the highly intermittent flow pattern found there in LES. DOI: 10.1115/1.3112386
Introduction
poral scales of the turbulent motion. With DNS, the complete flow
field throughout the domain is determined with no modeling assumptions, and thus it offers the most accurate characterization of
the turbulence possible. The principal drawback of DNS is that its
computational cost increases in proportion to the cube of the Reynolds number 2. This limits its use mainly to fundamental scientific investigations of turbulence, and makes it impractical for
CFD of the sort required for the present application.
At the other extreme in terms of computational cost is the
Reynolds-averaged NavierStokes RANS modeling. With
RANS, modeled equations for the mean velocity of a statistically
stationary turbulent flow are solved. The higher-order Reynolds
stresses are usually obtained from a turbulent viscosity model,
which is algebraically related to other turbulence quantities such
as the turbulent kinetic energy k and energy dissipation rate for
which modeled transport equations are solved. Because RANS is
not concerned with solving for the fluctuating velocity and pressure fields, the computational cost is much lower by orders of
magnitude than the equivalent DNS. However, the accuracy of
RANS predictions is limited by the fidelity of the closure models
used to represent the effects of the turbulent fluctuations. Despite
decades of research, there remain fundamental limitations to what
can be expected from a RANS simulation 3.
Large eddy simulation LES has emerged as an important intermediate approach, whereby one solves for the instantaneous
filtered velocity of the largest energy-containing eddies of the turbulent motion and models the effect of the subgrid-scale motions
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
In forced convection with low turbulence levels, the renormalization group RNG k- model by Yakhot and Orszag 25, the low
Reynolds number k- model due to Launder and Sharma LRN-LS
26, the v2f-dav, and LES all performed very well. The v2f
model by Durbin 27 was found to be very promising but the
numerical algorithm has to be constructed so as to avoid some
inherent numerical problems. The RSMs were shown to be able to
capture more flow details than eddy viscosity based models; however, they require more computational time. The authors concluded that LES provided more detailed and possibly more accurate predictions for indoor air flows and that it is a very important
tool for understanding the dynamics of indoor airflow.
In the current investigation, we use LES to model the air flow
inside an experimental flow chamber located in the Indoor-FlowLaboratory IFL at Syracuse University. The hydrodynamic conditions within the facility are typical of displacement ventilation
systems, where the supply air displaces the room air with presumed minimal mixing so as to achieve a high air change efficiency 1.
Consistent with the incremental procedure for simulation of displacement ventilation systems laid out by Chen and Srebric 28
and Chen and Zhai 29, we have neglected thermal convection
effects and geometrical complexities at this stage. The main objective here was limited to modeling and understanding the basic
flow behavior inside the IFL, with an eye toward incorporating
more comprehensive physics and more geometrical details in future investigations.
The sensitivity of the flow field to the inlet turbulence levels
and the details of the flow at the inlet are investigated. The performance of the Smagorinsky LES model with constant coefficient and the dynamic model in predicting the air flow are systematically studied and analyzed. Finally, light is shed on the
performance of the k- model in predicting the mean flow and the
turbulent kinetic energy throughout the flow field at varying inlet
turbulence levels.
Turbulence Models
ui
=0
+
t
xi
ui
ui u j 2 ul
uiu j =
+
ij
+
t
xj
xi x j
x j xi 3 xl
+
uiuj
xj
where is the density, p is the pressure, is the dynamic viscosity, ui is the velocity component in the xi direction, and the overbar indicates an averaged quantity. The symbol ij denotes the
Kronecker delta, and the Einstein summation convention is used.
The Reynolds-averaged approach to turbulence modeling requires
that the Reynolds stress term uiuj be appropriately modeled. In
the k- model, the Boussinesq approximation is used to relate the
Reynolds stresses to the mean velocity gradients through
uiuj = t
uk
ui u j
2
+
k + t
ij
3
xk
x j xi
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
i =
ku
+
t xi
xj
ui =
+
t xi
xj
t k
+ Gk
k x j
ij u
uiu j
iu j
ij 31 kkij = 2tSij
11
12
where Sij is the rate-of-strain tensor for the filtered velocity, defined by
S 1 ui + u j
ij
2 x j xi
13
t = Ls2S
14
Ls = mind,CsV1/3
15
ui
ij
p ij
uiu j =
+
t
xj
xj xj
xi x j
10
ui
=0
+
t
xi
2 ul
ui u j
+
ij
3 xl
x j xi
The subgrid stresses are modeled using the Boussinesq approximation as in the RANS models:
5
k
=0
n
that represents the contributions of the filtered velocity to the viscous stress tensor, and ij is the subgrid-scale stress defined by
2
t
+ C 1 G k C 2
k
k
x j
k2
t = C
ij =
8
9
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
244
15
20
Z
Y
Outlet
46
Inlet
183
(a)
(b)
Fig. 1 a Geometry of Indoor Flowfield Laboratory Chamber; dimensions are 244 183 244 cm3.
b Top view; dimensions are in cm.
edness criterion is violated 30. Provided that the numerical solution converges, this approach leads to pure second-order differencing 30. The bounded central difference scheme was used for
discretization of both the pressure and momentum equations, and
default tolerances were used a residual tolerance of 0.0001 and
relaxation factors of 1 for the pressure and momentum equations.
Air with constant density and viscosity at standard sea level conditions was used. The time step was set so that the resulting maximum Courant number would be no larger than 0.5. The boundary
conditions used in the first stage of this investigation were simple
plug/uniform flow velocity profile at the inlet, with uniform turbulence intensities of 0%, 5%, and 13%, a uniform turbulence
length scale equal to 7% of the inlet hydraulic diameter, a pressure
outlet for the outflow, and solid wall boundary conditions everywhere else. Simulations of the flow were continued up to the time
when the flow was approximately statistically stationary. In the
second stage of the study, the inlet velocity and turbulence profiles
obtained from experiments conducted in the IFL at Syracuse University 35 were used in the simulations.
3.1 Description of the Geometry. The IFL chamber consists
of a 2.44 1.83 2.44 m3 cubicle, with optical access for particle
image velocimetry PIV measurements through the front 1.83
2.44 m2 wall. The origin of the coordinate system is located at
the center of the cubicle, with positive z pointing vertically upward and positive x pointing toward the front wall see Fig. 1.
The chamber half-height L is used to nondimensionalize the vertical distance. The chamber is designed to represent a typical
indoor-flow environment for a single occupancy cubicle. An isometric view of the chamber is shown in Fig. 1a and a top view
is shown in Fig. 1b. The cubicle is ventilated by a low speed,
closed-loop system that is designed to allow an occupants thermal plume to be a significant factor in driving the flow. There is a
0.2 0.46 m2 inlet on the floor, near the front wall, and an identically sized outlet in the ceiling near the rear wall. The average
inlet and outlet flow velocity is 0.2 m/s.
3.2 Description of the Computational Grid. An initial grid
containing 48 32 48 cells was developed using FLUENTs Gambit mesh generator to represent the chamber. The accuracy of the
resulting solution was checked by grid refinement and comparison
of the resulting solution with the original solution on the coarser
grid. The process was repeated until an accurate solution was
obtained. Based on the initial results from the coarse grid, the grid
was locally refined in those regions exhibiting steep gradients of
the solution variables such as the inlet, the shear layers around the
inflow plume, the ceiling, and the outlet. First, a vertical column
051402-4 / Vol. 131, MAY 2009
including the inlet and three cells on each side of it was refined
once by halving the mesh dimensions in each coordinate direction, and another volume around the outlet, extending three cells
on each side of it and four cells below it was refined again, by
halving the mesh dimensions in each coordinate direction. Finally, a volume covering the entire ceiling area and extending four
cells below, it was similarly refined. The resulting grid was again
used to determine the general features of the flow. Based on the
results of the initial calculations on this grid, it was determined
that extra refinement of the grid in the vicinity of the inlet was
required to maintain adequate accuracy. Two additional steps of
refinement were done. The first involved refinement of the volume
covering the inlet area and four cells away from it on all sides, and
extending 18 cells above the inlet. The second refinement covered
the area of the inlet and two cells away from it on all sides, and
extended nine cells above the inlet. The resulting grid, referred to
here as the baseline grid, is shown in Figs. 2a and 2b where a
shows a side view of the vertical plane passing through the center
of the inlet and b shows a top view of the floor and the inlet area.
The baseline grid calculations were performed on a Linux cluster
using ten processors. The wall-clock time per time step was about
10 s corresponding to 800 s of wall-clock compute time per
physical second since the time step on this grid is t = 0.0125 s.
It was found that the wall-clock time for calculations with the
dynamic subgrid-scale model is nearly the same as that for the
constant coefficient Smagorinsky model. To test grid convergence,
one case 13% inlet turbulence intensity was solved on a refined
grid that contained twice the number of grid cells in each direction. The time step used on the refined grid was t = 0.006 25 s,
so the solution for a fixed time interval was approximately 16
times more computationally expensive than the solution on the
baseline grid. The computations on the refined grid were performed on a newer Linux cluster utilizing Infiniband as an interconnect, which was found to significantly decrease the wall-clock
time. Using 20 cores on five nodes each having two dual core
processors reduced the wall-clock time to about 1 s per time step.
3.3 Simulation of Inlet Turbulence. Simulation of inlet turbulence for fully developed turbulent flows can, in principle, be
achieved by adding small perturbations to a laminar flow and by
having a long enough computational domain in the streamwise
direction to allow the turbulence to develop. In order to reduce the
development region, a more viable approach is to introduce an
upstream domain and to solve for the flow assuming periodicity in
the streamwise direction 36. The result of this calculation is then
used as an inflow boundary condition for the main simulation.
Transactions of the ASME
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Y
(a)
Z
(b)
Fig. 2 Computational grid for the Indoor Flowfield Laboratory Chamber: a side x-z plane and b top x-y plane
views of the baseline grid
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Refined grid
Baseline grid
0.2
0.08
0.175
0.15
0.125
0.1
0.075
0.05
0.025
(a)
0
0.7
Refined grid
Baseline grid
0.09
0.07
0.06
0.05
0.04
0.03
0.02
0.01
0.8
0.9
1.1
1.2
(b)
x [m]
0
0.7
0.8
0.9
1.1
1.2
x [m]
Fig. 3 Comparison of mean velocity magnitude on the baseline and the refined grids at a z / L = 0.2 and b z / L
= 0.9375
energy. If this is achieved, LES should be at most weakly dependent on the particular subgrid-scale model used in the computations. In FLUENT, the mixing length for the subgrid scales Ls and
the subgrid-scale eddy viscosity t = t / are used to construct an
estimate for the subgrid-scale turbulent kinetic energy ks defined
as
ks =
t2
16
Ls2
on the refined grid the subtest kinetic energy is less than 5% of the
total turbulent kinetic energy, except very near the inlet, so the
solution on this grid is judged to be well resolved. The subtest
kinetic energy on the baseline grid is somewhat larger, but still
small enough relative to the total turbulent kinetic energy that we
judge the solution on the baseline grid to be adequately resolved
for the subsequent computations. This conclusion is supported by
the earlier comparison of the mean and rms profiles shown in
Figs. 3 and 4.
4.2 Effect of Inlet Turbulence Level Using Plug-Flow Inlet
Conditions. Figures 6a and 6b show the contour plots of the
instantaneous and mean velocity magnitudes, respectively, along
the vertical x-z plane passing through the middle of the inlet for
the 0% inlet turbulence intensity case. It is clear that the jet in this
case is very confined and that there is little mixing with the room
air. The jet flows straight up toward the ceiling, bends 90 deg, and
flows along the ceiling until it reaches the exit. Other than this
basic flow pattern, there are no significant organized motions or
Refined grid
Baseline grid
0.08
0.07
0.07
0.06
0.05
0.04
0.03
0.02
0.01
(a)
0
0.7
Refined grid
Baseline grid
0.08
0.06
0.05
0.04
0.03
0.02
0.01
0.8
0.9
x [m]
1.1
1.2
(b)
0
0.7
0.8
0.9
1.1
1.2
x [m]
Fig. 4 Comparison of the rms of the velocity magnitude on the baseline and the refined grids at a z / L = 0.2 and b
z / L = 0.9375
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
0.003
0.002
0.0025
0.0015
T.K.E. [m /s ]
T.K.E. [m /s ]
0.002
0.0015
0.001
0.001
0.0005
0.0005
0
0.7
0.8
0.9
(a)
1.1
0
0.7
1.2
x [m]
(b)
0.8
0.9
1.1
1.2
x [m]
Fig. 5 Comparison of turbulent kinetic energy on the baseline and the refined grids at a z / L = 0.8 and b z / L
= 0.9375
0.258
0.2451
0.2322
0.2193
0.2064
0.1935
0.1806
0.1677
0.1548
0.1419
0.129
0.1161
0.1032
0.0903
0.0774
0.0645
0.0516
0.0387
0.0258
0.0129
0
0.281
0.26695
0.2529
0.23885
0.2248
0.21075
0.1967
0.18265
0.1686
0.15455
0.1405
0.12645
0.1124
0.09835
0.0843
0.07025
0.0562
0.04215
0.0281
0.01405
0
(a)
(b)
Fig. 6
a Instantaneous and b mean velocity contours on center plane; 0% inlet turbulence units are in m/s
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
0.262
0.2489
0.2358
0.2227
0.2096
0.1965
0.1834
0.1703
0.1572
0.1441
0.131
0.1179
0.1048
0.0917
0.0786
0.0655
0.0524
0.0393
0.0262
0.0131
0
0.274
0.2603
0.2466
0.2329
0.2192
0.2055
0.1918
0.1781
0.1644
0.1507
0.137
0.1233
0.1096
0.0959
0.0822
0.0685
0.0548
0.0411
0.0274
0.0137
0
(a)
(b)
Fig. 7
a Instantaneous and b mean velocity contours on center plane; 5% inlet turbulence units are in m/s
0.262
0.2489
0.2358
0.2227
0.2096
0.1965
0.1834
0.1703
0.1572
0.1441
0.131
0.1179
0.1048
0.0917
0.0786
0.0655
0.0524
0.0393
0.0262
0.0131
0
0.304
0.2888
0.2736
0.2584
0.2432
0.228
0.2128
0.1976
0.1824
0.1672
0.152
0.1368
0.1216
0.1064
0.0912
0.076
0.0608
0.0456
0.0304
0.0152
0
(a)
(b)
Fig. 8
a Instantaneous and b mean velocity contours on center plane; 13% inlet turbulence units are in m/s
Laminar inflow
5% Turbulence
13% Turbulence
0.2
0.175
0.175
0.15
0.125
0.1
0.075
0.05
0.025
(a)
0.7
Laminar inflow
5% Turbulence
13% Turbulence
0.2
0.15
0.125
0.1
0.075
0.05
0.025
0.8
0.9
x [m]
1.1
1.2
(b)
0.7
0.8
0.9
1.1
1.2
x [m]
Fig. 9 Comparison of the mean velocity magnitude for three levels of inlet turbulence intensity at a z / L = 0.4 and b
z / L = 0.0; baseline grid
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Laminar inflow
5% Turbulence
13% Turbulence
0.16
0.14
0.12
0.1
0.08
0.06
0.04
0.02
0
0.7
0.8
0.9
1.1
1.2
x [m]
Profiles of the velocity rms are shown in Figs. 11 and 12. Once
again the 5% and 13% cases show quite similar profiles throughout the computational domain. The behavior of the laminar inflow
case needs some elaboration. At the z / L = 0.4 and z / L = 0.0 stations, the level of turbulence is significantly less than for the 5%
and 13% cases because of the laminar state of the entering jet.
However, near the ceiling, the shear layers have grown enough to
begin to interact with the front wall and turbulence is generated
there and also in the ceilings boundary layer. The contours in Fig.
6 show the generation of a blob of high velocity fluid near the
ceiling, which can contribute to the high levels of turbulence near
the ceiling. Accordingly, the rms levels of the velocity fluctuations
in the laminar inflow case are seen to be higher than the rms levels
for the 5% and 13% cases near the ceiling see Figs. 11 and 12.
The significant differences between the simulation results for
the 0% and both the 5% and 13% cases cannot be captured by the
standard k- model, which is designed primarily for fully developed turbulent flows. This very fact led Loomans 42, who investigated the effect of inlet turbulence intensity level on the flow
field of a displacement ventilation system in a full-scale room
using the standard k- model, to conclude that the level of inlet
Laminar inflow
5% Turbulence
13% Turbulence
0.06
0.05
0.05
0.04
0.03
0.02
0.04
0.03
0.02
0.01
(a)
0
0.7
Laminar inflow
5% Turbulence
13% Turbulence
0.06
0.01
0.8
0.9
1.1
1.2
(b)
x [m]
0
0.7
0.8
0.9
1.1
1.2
x [m]
Fig. 11 Comparison of the rms velocity magnitude for three levels of inlet turbulence intensity at a z / L = 0.4 and b
z / L = 0.0; baseline grid
0.065
Laminar inflow
5% Turbulence
13% Turbulence
0.08
0.07
0.06
0.05
0.04
0.7
0.8
0.9
(a)
x [m]
1.1
Laminar inflow
5% Turbulence
13% Turbulence
0.06
0.055
0.05
0.045
0.04
0.7
1.2
(b)
0.8
0.9
1.1
1.2
x [m]
Fig. 12 Comparison of the rms velocity magnitude for three levels of inlet turbulence intensity at a z / L = 0.875 and b
z / L = 0.9375; baseline grid
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
0.25
0.3
0.25
0.2
0.2
Exp.Data/Left side
0.15
Exp.Data/Right side
W [m/s]
W [m/s]
0.15
0.1
0.05
Experimental data
0.05
0.05
1
(a)
0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
Y/L
(b)
0
800
850
900
950
1000
1050
1100
X [mm]
Fig. 13 Experimental data and curve fits along the a long and b short sides of the inlet. In a, Ls is the half-length of
the inlet section.
the front wall, due to distortions and reflections from the glass
wall; consequently, assumptions had to be made about the profiles
of the measured quantities. The experimental data consist of the
mean values of the three components of velocity and the rms of
their fluctuations. The mean velocity components in the plane of
the inlet, i.e., the U and V components, were deemed small
enough to neglect, whereas the vertical component W of the velocity was fitted to a smooth polynomial. The variations parallel to
the long side of the inlet were not as significant as those across the
short side, and a fit of the averaged experimental data on both
sides was used. Figure 13a shows two fourth-order polynomials
one covering most of the profile and one representing the shear
layer portion. In the direction of the short side, the interpretation
of the data was complicated by the fact that part of the inlet was
not covered by the measurement window; hence this part of the
data was missing. We extrapolated the missing data using two,
fourth-order, polynomials that represent the profile, including the
missing segment, as shown in Fig. 13b. The turbulent kinetic
energy and the root mean square of the fluctuating velocities were
assumed symmetric across the short side of the duct, with negligible variation in the direction of the long side, as shown in Fig.
14 which shows only half of the profile since it is assumed
symmetric.
FLUENT calculations with these fitted inlet profiles were run and
compared with the 13% case with uniform profiles. Figures 15
and 16 show comparisons of the mean velocity at four vertical
stations. Note that the significant differences at the lowest station
z / L = 0.8 become weaker with increasing distance from the inlet. The maximum deviation between the two cases occurs in the
shear layers, where the deviation of the measured inlet turbulent
kinetic energy from the spatially averaged value is greatest. This
suggests that there is some memory of the inlet conditions
throughout the flow.
Similar behavior of the rms of the velocity is shown in Figs. 17
and 18, where close to the inlet at the z / L = 0.8 and z / L = 0.4
stations the profiles are significantly different, whereas close to
the ceiling, the two cases predict similar levels of turbulence notice the offset of the origin of the vertical axis in Fig. 18 amplifies
the small differences between the profiles. So in general the flow
field exhibits sensitivity to the inlet flow conditions primarily near
the inlet. As the flow evolves spatially, the effect of the inlet
conditions becomes less pronounced. This conclusion is in agreement with Jiangs 44 conclusions from his LES and k- model
calculations for similar comparisons. Joubert et al. 43 also reported that using a parabolic inlet velocity profile as opposed to a
Transactions of the ASME
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
x 10
Experimental data
Second degree polynomial fit
0
860
880
900
920
940
960
980
X [mm]
uniform profile does not affect the mean velocity in the twodimensional cavity flow that they analyzed with the k- model;
however, they found significant differences in the predicted levels
of turbulence in the cavity, which they explained on the basis of
the contribution of the shearing already present in the inflow in
the parabolic profile case. It should be noted that the wall jet
studied in their two-dimensional simulation is fundamentally different from the present configuration. The close proximity of the
jet to the top boundary may have caused the greater sensitivity.
They also noted that the width of the slot has an important effect
on the turbulence level within the cavity.
4.4 Comparison of the Constant Coefficient and Dynamic
Smagorinsky LES Models. The success of a subgrid-scale model
can be characterized by how well it predicts the large-scale statis-
0.15
0.1
0.05
0
0.7
(a)
0.2
0.2
0.8
0.9
x [m]
1.1
0.15
0.1
0.05
0
0.7
1.2
(b)
0.8
0.9
1.1
1.2
x [m]
Fig. 15 Mean velocity magnitude at a z / L = 0.8 and b z / L = 0.0; comparison of results with plug flow and experimentally determined inlet conditions
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
0.1
0.14
0.12
0.1
0.08
0.06
0.04
0.08
0.06
0.04
0.02
0
0.7
0.8
0.9
1.1
0.7
1.2
0.8
0.9
1.1
1.2
(a)
(b)
x [m]
x [m]
Fig. 16 Mean velocity magnitude at a z / L = 0.75 and b z / L = 0.875; comparison of results with plug flow and experimentally determined inlet conditions
0.05
0.05
0.045
0.045
0.04
0.035
0.03
0.025
0.02
0.015
0.01
(a)
0.7
0.04
0.035
0.03
0.025
0.02
0.015
0.01
0.8
0.9
1.1
1.2
0.7
0.8
0.9
1.1
1.2
(b)
x [m]
x [m]
Fig. 17 rms velocity magnitude at a z / L = 0.8 and b z / L = 0.4; comparison of results with plug flow and experimentally determined inlet conditions.
0.065
0.06
0.055
0.05
0.045
0.04
0.7
0.8
0.9
0.065
0.07
1.1
1.2
0.06
0.055
0.05
0.045
0.04
0.7
0.8
0.9
1.1
1.2
(b)
x [m]
x [m]
Fig. 18 rms velocity magnitude at a z / L = 0.875 and b z / L = 0.9375; comparison of results with plug flow and experimentally determined inlet conditions
(a)
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
0.2
0.15
0.1
0.05
0
0.7
0.8
0.9
1.1
0.15
0.1
0.05
0
0.7
1.2
0.8
0.9
1.1
1.2
(b)
x [m]
x [m]
Fig. 19 Mean velocity magnitude at a z / L = 0.4 and b z / L = 0.0; comparison of results using constant-coefficient and
dynamic Smagorinsky models
(a)
0.08
0.1
0.08
0.06
0.04
0.02
0.7
0.8
0.9
1.1
0.07
0.06
0.05
0.04
0.7
1.2
0.8
0.9
1.1
1.2
(b)
x [m]
x [m]
Fig. 20 Mean velocity magnitude at a z / L = 0.875 and b z / L = 0.9375; comparison of results using constant-coefficient
and dynamic Smagorinsky models
(a)
0.05
0.06
0.045
0.055
0.04
0.035
0.03
0.025
0.7
0.045
0.04
0.035
0.03
0.02
(a)
0.05
0.8
0.9
1.1
1.2
0.7
0.8
0.9
1.1
1.2
(b)
x [m]
x [m]
Fig. 21 rms velocity magnitude at a z / L = 0.4 and b z / L = 0.0; comparison of results using constant-coefficient and
dynamic Smagorinsky models
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
0.07
0.06
0.065
0.06
0.055
0.05
0.055
0.05
0.045
0.045
0.7
(a)
0.8
0.9
1.1
1.2
x [m]
(b)
0.04
0.7
0.8
0.9
1.1
1.2
x [m]
Fig. 22 rms velocity magnitude at a z / L = 0.875 and b z / L = 0.9375; comparison of results using constant-coefficient
and dynamic Smagorinsky models
k- model, 0% turbulence
k- model, 13% turbulence
k- model, 0% turbulence
k- model, 13% turbulence
0.08
0.2
0.15
0.1
0.05
0
0.7
(a)
0.8
0.9
x [m]
1.1
0.06
0.04
0.02
1.2
(b)
0.8
0.9
1.1
1.2
x [m]
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
k- model, 0% turbulence
k- model, 13% turbulence
k- model, 0% turbulence
k- model, 13% turbulence
0.0012
0.0025
0.001
0.0008
T.K.E. [m /s ]
T.K.E. [m /s ]
0.002
0.0015
0.001
0.0006
0.0004
0.0005
0.0002
0.7
(a)
0.8
0.9
1.1
1.2
x [m]
0.7
(b)
0.8
0.9
1.1
1.2
x [m]
of inlet turbulence especially far from the inlet. The slow development of the jet predicted by LES in the 0% case is not observed
with the k- model. To emphasize this point, we compare the
solution from the k- and LES, using the constant coefficient
model, at 0% and 13% inlet turbulent levels. Figure 25 shows the
comparison for the mean velocity magnitude profile very close to
the inlet z / L = 0.8 at 0% and 13% inlet turbulent intensities. At
0% turbulence intensity, LES is predicting a flat velocity profile
with a sharp gradient separating the potential core from the quiescent fluid outside. The k- model, in contrast, predicts a much
more diffuse profile with curved shape similar to the profile for a
fully developed free jet. The two model predictions are in better
agreement at the 13% turbulence intensity, as the turbulence is
more nearly fully developed leading both models to predict a
smoother profile. The inability of the k- model to predict the
slow development of the jet at 0% inlet turbulence intensity leads
to an overprediction of the spreading rate. As discussed in Ref.
2, a well-known deficiency of the k- model is that it signifi-
cantly overpredicts the rate of spreading for the round jet. Although these differences could be reduced by adjusting the value
of C1 or C2, we did not feel that this would be justified. Figure
26 shows the mean velocity profile at a location well removed
from the inlet z / L = 0.75. The agreement between the LES and
k- predictions is better at the 13% inlet turbulence intensity level.
Nevertheless, it is clear that the k- model fails to capture the
correct qualitative behavior near the upper right corner x
1.22 m, as it shows a nearly stagnant region, whereas the LES
model predicts that there are small patches of intermittent activity
in the corner. Figure 27 shows a comparison of the resolved turbulent kinetic energy from the LES model to the turbulent kinetic
energy from the k- model at 0% and 13% inlet turbulent intensities. In the laminar inflow case, it is clear that LES is predicting
the correct physical behavior of the jet close to the inlet where
there is no active mechanism for turbulence generation yet at
z / L = 0.8 as the shear layer thickness is close to zero and the jet
0.2
0.15
0.1
0.05
0
0.7
(a)
0.2
0.8
0.9
x [m]
1.1
0.15
0.1
0.05
0
0.7
1.2
(b)
0.8
0.9
1.1
1.2
x [m]
Fig. 25 Mean velocity magnitude from k- and LES models at z / L = 0.8: a 0% and b 13% inlet turbulence intensities
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
0.15
0.1
0.05
0
0.7
0.8
0.9
1.1
0.1
0.05
0
0.7
1.2
x [m]
(a)
0.15
0.8
0.9
1.1
1.2
x [m]
(b)
Fig. 26 Mean velocity magnitude from k- and LES models at z / L = 0.75: a 0% and b 13% inlet turbulence intensities
Conclusions
with 0%, 5%, and 13% inlet turbulence intensities. It is seen that
the 5% and 13% cases exhibit much faster spreading and mixing
with the room air than the laminar inflow case and that the statistics predicted in these two cases are close to each other, especially
far from the inlet. We conjecture that beyond a threshold level of
inlet turbulence intensity, the jet develops nearly independently of
the inlet turbulence intensity. This is consistent with the findings
of Jiang 44, Loomans 42, and Joubert et al. 43 who likewise
found little sensitivity to the inlet turbulence levels. However, the
laminar inlet flow case yielded significant deviations of the mean
flow and turbulence levels from the simulations with higher levels
of inlet turbulence. The results of flow simulations with experimentally measured profiles for the inlet conditions are compared
with those for the 13% case with uniform mean inlet profiles. It is
seen that near the inlet the flow is sensitive to the inlet flow
details, but these effects become less pronounced, although nonnegligible, further away from the inlet. The results of simulations
using the constant coefficient Smagorinsky and the dynamic LES
0.0035
0.003
0.003
0.0025
T.K.E. [m /s ]
2
2
T.K.E. [m /s ]
0.0025
0.002
0.0015
0.001
0.7
0.0015
0.001
0.0005
(a)
0.002
0.0005
0.8
0.9
x [m]
1.1
1.2
(b)
0.7
0.8
0.9
1.1
1.2
x [m]
Fig. 27 Turbulent kinetic energy from k- and LES models at z / L = 0.8: a 0% and b 13% inlet turbulence intensities
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
0.003
0.0025
0.0025
2
T.K.E. [m /s ]
0.002
0.003
T.K.E. [m /s ]
0.0015
0.0005
0.0005
(a)
0.0015
0.001
0.001
0.7
0.002
0.8
0.9
1.1
1.2
x [m]
(b)
0.7
0.8
0.9
1.1
1.2
x [m]
Fig. 28 Turbulent kinetic energy from k- and LES models at z / L = 0.75: a 0% and b 13% inlet turbulence intensities
models agree reasonably well especially far from the inlet where
most of the turbulent kinetic energy is resolved. The most significant differences are seen near the inlet, where the subgrid-scales
contribution to the total turbulent kinetic energy is as high as 30%
of the total turbulent kinetic energy. The k- model simulations
are compared at 0% and 13% inlet turbulence intensities. It is seen
that the k- model is nearly insensitive to the level of inlet turbulence especially far from the inlet. Finally, the results of simulations using the k- and LES models with 0% and 13% inlet turbulence intensities are compared. It is seen that the k- model
predictions for the 0% case are significantly different than the
LES predictions, as the k- model fails to capture the slow spatial
development of the jet into a fully turbulent state. At the 13%
turbulence level, it is seen that the k- model predicts the general
features of the mean velocity reasonably well. Even for this case,
the k- overpredicts the spreading rate of the jet relative to the
LES model and fails to capture the complicated unsteady flow
pattern near the ceiling leading to significant overprediction of the
turbulent kinetic near the inlet and underprediction of it near the
ceiling.
Acknowledgment
This work was supported in part by the Environmental Protection Agency, through the Syracuse University NY STAR Center
for Environmental Quality Systems/EPA Indoor Environmental
Research Program Collaboration.
References
1 Spengler, J. J., Samuel, J. M., and McCarthy, J. F., 2001, Indoor Air Quality
Handbook, McGraw-Hill, New York.
2 Pope, S. B., 2000, Turbulent Flows, Cambridge University Press, Cambridge.
3 Wilcox, D. C., 2004, Turbulence Modeling for CFD, DCW Industries, Inc., La
Canada, CA.
4 Smagorinsky, J., 1963, General Circulation Experiments With the Primitive
Equations: I. The Basic Equations, Mon. Weather Rev., 91, pp. 99164.
5 Lilly, D. K., 1967, The Representation of Small-Scale Turbulence in Numerical Simulation Experiments, Proceedings of the IBM Scientific Computing
Symposium on Environmental Sciences, IBM Form 320-1951, pp. 195210.
6 Deardorff, J. W., 1974, Three-Dimensional Numerical Study of the Height
and Mean Structure of a Heated Planetary Boundary Layer, Boundary-Layer
Meteorol., 7, pp. 81106.
7 Schumann, U., 1975, Subgrid-Scale Model for Finite Difference Simulations
of Turbulent Flows in Plane Channels and Annuli, J. Comput. Phys., 18, pp.
376404.
8 Pope, S. B., 2004, Ten Questions Concerning the Large-Eddy Simulation of
Turbulent Flows, New J. Phys., 635, pp. 124.
9 Germano, M., Piomelli, U., Moin, P., and Cabot, W. H., 1991, A Dynamic
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
39 Lee, S., Lele, S., and Moin, P., 1992, Simulation of Spatially Evolving Turbulence and the Applicability of Taylors Hypothesis in Compressible Flow,
Phys. Fluids A, 4, pp. 15211530.
40 Smirnov, A., Shi, S., and Celik, I., 2001, Random Flow Generation Technique
for Large Eddy Simulation and Particle Dynamics Modeling, ASME J. Fluids
Eng., 123, pp. 359371.
41 Mathey, F., Cokljat, D., Bertoglio, J. P., and Sergent, E., 2003, Specification
of LES Inlet Boundary Condition Using Vortex Method, Fourth International
Symposium on Turbulence, Heat and Mass Transfer, K. Hanjalic, Y. Nagano,
and M. Tummers, eds., Begell House, Inc., Antalya, Turkey.
42 Loomans, M., 1998, The Measurement and Simulation of Indoor Air Flow,
Ph.D. thesis, Eindhoven University of Technology, Eindhoven, The Netherlands.
43 Joubert, P., Sandu, A., Beghein, C., and Allard, F., 1996, Numerical Study of
the Influence of Inlet Boundary Conditions on the Air Movement in a Ventilated Enclosure, Proceedings of the Roomvent, Yokohama, Japan, Vol. 1, pp.
235242.
44 Jiang, J., 2007, Experimental and Numerical Study of Air Flows in a FullScale Room, Ph.D. thesis, University of Illinois at Urbana-Champaign, Urbana.
45 Marr, D., 2007, private communication.
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
H. Zidouh
L. Labraga1
e-mail: llabraga@univ-valenciennes.fr
M. William-Louis
Universit Lille Nord de France,
F-59000 Lille, France;
UVHC, LME,
59313 Valenciennes, France
Introduction
wt =
tVtVt
2
where wut is zero for steady flows, small for slow transients,
and significant for fast transient flows. Daily et al. 1 conducted
laboratory experiments and found wut to be positive for accelerating flows and negative for decelerating flows. They argued
that during acceleration, the central portion of the stream moved
somehow bodily so that the velocity profile steepened, giving
1
Corresponding author.
Contributed by the Fluids Engineering Division of ASME for publication in the
JOURNAL OF FLUIDS ENGINEERING. Manuscript received August 18, 2008; final manuscript received February 24, 2009; published online April 15, 2009. Assoc. Editor:
Hassan Peerhossaini.
w = qs +
kD dV
8 dt
wt = qst + wut
w = qs + k3
V
R V
a
x
2 t
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Theory
Shear stress is determined experimentally using an electrochemical method based on the reduction in the ferricyanide ion on
an electrode surface. A thorough review of this technique is provided by Hanratty and Campbell 7. This method is based on the
determination of the limiting diffusion current at the surface of an
electrode under conditions for which the chemical reaction rate is
fast enough so that the concentration of the reacting ions is zero at
the surface of the working electrode. If the length of this electrode
is very small in the flow direction, the concentration boundary
layer on the electrode is thin; thus the flow velocity varies linearly
throughout the thickness of this boundary layer.
A diagram of a single electrochemical probe is given in Fig. 1.
The current I flowing to the test electrode of area A is related to
the mass transfer coefficient K by
K=
I
AneFC0
K
= cS+1/3
D
2 Sqt
Sct = Sqt + t0
3
t
where t0 = 0.4862/3D1/3S2/3
and Sqt = D / 2K+t / 0.8073.
q
This method is limited both to rather moderate frequencies and
nonreverse unsteady flows. In the presence of large unsteady or
reverse flows, the inverse method 1315 seems to be more appropriate to calculate the wall shear rate from the mass transfer
signal measured. This technique consists in solving the direct
convection-diffusion equation and in estimating sequentially the
unknown wall shear rate by minimizing the difference between
measured and simulated mean concentration gradients on the
probe surface. In this study both the inverse method and that
suggested by Sobolik et al. 12 were applied for the determination of the true wall shear rate of the transient turbulent flow in
a pipe.
The ultrasonic pulsed Doppler velocimetry 1618 was used
for velocity profile measurements. Figure 2 shows a common configuration of the ultrasound beam and flow. This method is based
on detecting and processing the backscattered echoes originating
from moving targets suspended in the flowing liquid that is to be
investigated. This technique provides measurements of local onedimensional velocity and related distance from the transducer,
leading to the construction of an almost instantaneous velocity
profile along the acoustic beam. Information on the position from
which the ultrasound is reflected is extracted from the time delay
TD between the start of the pulse burst and its reception as Y
= aTD / 2.
If the particle is moving at an angle regarding the axis of the
ultrasonic beam, its velocity u is measured by computing the
variation of its depth between two emissions separated by the time
interval between two pulses as
u=
fD
a
2f cos
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
1024
1.025 106
7.45 1010
FeCN3
6 + e FeCN6
f D max f prf/2
a
4Tprf f cos
10
4
10
5
68
0.69
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Fig. 5 Experimental results showing pressure time history and associated velocity profiles
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Kw p
Kw D
1 + 1 p
E e
11
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
q =
64
laminar
Re
0.316
turbulent
R0.25
e
12
The results show that the wall shear stress measured from the
electrochemical signals oscillates around that predicted by a quasisteady model for the time range between t = 0 s, corresponding
to the start of the valve closure, up to t = 0.055 s. The ability of
the electrochemical method to sense the local instantaneous flow
unsteadiness is clearly shown in Fig. 9. Indeed, the wall shear
stress exhibits a wavy behavior probably due to the pressure wave
originating from the valve closure. As for the quasisteady model,
it tends to smoothen out the actual wall shear stress fluctuations.
Shuy 2 found that the unsteady wall shear stress is higher than
the quasisteady values computed from the KarmanNikuradse
equation based on instantaneous velocity, for a value of the acceleration parameter = 2D / sV2dV / dt up to 2. The values of
the acceleration parameter of the present study range within 0
100 for t 0.05 s. For this severe decelerating flow, the
present results show that the unsteady wall shear stress may be
either greater or lower than that obtained from a quasisteady approach, following the local fluctuating acceleration with a phase
shift, as it is shown below.
For 0.06 s t 0.07 s, the quasisteady assumption is no
longer valid to determine the wall shear stress. The effective wall
velocity gradient is considerably higher than the one resulting
from the pseudosteady approach. The unsteady wall shear stress
inferred from the electrochemical probe increases up to a maximum value corresponding to that of the pressure. It is shown that
the maximum unsteady wall shear stress value reaches the steady
one at the end of this phase.
For t 0.07 s, the unsteady wall shear stress tends asymptotically toward zero with large oscillations resulting from all reflected pressure waves.
Figure 10 gives a synthesis of the results of both unsteady wall
shear stress component and acceleration wu. According to Eq.
3, it is expected that the changes in the unsteady wall shear
stress component wu correspond closely to those of the acceleration. It is shown that the sudden change in the core of the pipe first
affects the acceleration before unsteady wall shear stress is
changed. The change in flow rate is then transferred toward the
wall by viscous shear. As a result, changes in the velocity field of
the fluid adjacent to the pipe wall, and in resulting wall shear
stresses, lag behind flow rate changes in the core of the pipe 22.
This should explain the discrepancies in phase shift between experimental and numerical head traces obtained in some previous
studies with the friction model based on instantaneous local acceleration for later time of the transient flows.
051403-6 / Vol. 131, MAY 2009
wall
shear
stress
component
and
Axworthy et al. 23 suggested that the unsteady friction formulas based on instantaneous accelerations such as Eq. 3 are
applicable to transient flow problems in which the wave passage
time-scale L / a = 0.0052 s in the present study is significantly
shorter than the vorticity diffusion time-scale 2D / 2u = 0.46 s
in the present study. According to Axworthy et al. 23, the
changes in wall shear stress in the conditions mentioned above
should correspond closely to the acceleration. In the light of the
present study, it can be stated that the time-scale arguments do not
allow any definitive conclusion to be made to clarify the applicability of the instantaneous-acceleration approach in both attenuation and phase shift of the transient wall shear rate.
Shuy 2 rewrote Eq. 3 in the following dimensionless form:
k
w
=1+
ws
2
13
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Nomenclature
a
A
C
d
D
F
f
Fig. 12 Dependence of k on the acceleration parameter
the short length of the pipe and the relatively high Reynolds number. The electrochemical technique is actually the only method
able to measure with a fine time resolution the unsteady wall shear
stress in these severe experimental conditions, not found in literature.
The computed values of the empirical coefficient k, obtained
from Eq. 13, are displayed against in Fig. 12. The values of k
are consistent with the distribution of the ratio between unsteady
and steady wall shear stresses analyzed above. The present results
show that the values of k are rather low compared with those
found in literature. Moreover, k is variable in time and therefore
depends on the acceleration parameter as it was already suggested
by Axworthy et al. 23.
Further experiments are performed in order to check the validity of the well known model of Brunone 4 Eq. 4, the most
widely used due to its ability to produce reasonable agreement
with experimental pressure history.
Conclusion
Fprf
I
k
K
ne
p
R
Re
S
t
Tprf
V
Special Characters
friction factor
kinematic viscosity m2 / s
fluid density kg/ m3
w wall shear stress Pa
ws quasisteady wall shear stress Pa
wu unsteady wall shear stress Pa
References
1 Daily, J. W., Hankey, W. L., Olive, R. W., and Jordan, J. M., 1956, Resistance
Coefficient for Accelerated and Decelerated Flows Through Smooth Tubes and
Orifices, Trans. ASME, 78, pp. 10711077.
2 Shuy, E. B., 1996, Wall Shear Stress in Accelerating and Decelerating Turbulent Pipe Flows, J. Hydraul. Res., 342, pp. 173183.
3 Vardy, A., and Brown, J. M. B., 1997, Discussion on Wall Shear Stress in
Accelerating and Decelerating Pipe Flows, J. Hydraul. Res., 351, pp. 137
139.
4 Brunone, B., Golia, U. M., and Greco, M., 1991, Some Remarks on the
Momentum Equation for Transients, International Meeting on Hydraulic
Transients With Column Separation, Ninth Round Table, IAHR, Valencia,
Spain, pp. 201209.
5 Carstens, M. R., and Roller, J. E., 1959, Boundary-Shear Stress in Unsteady
Turbulent Pipe Flow, ASCE J. Hydr. Div., 85HY2, pp. 6781.
6 Kurokawa, J., and Morikawa, M., 1986, Accelerated and Decelerated Flows
in a Circular Pipe 1st. Report, Velocity Profiles and Friction Coefficient,
Bull. JSME, 29249, pp. 758765.
7 Hanratty, T. J., and Campbell, J. A., 1983, Measurement of Wall Shear
Stress, Fluid Mechanics Measurements, R. J. Goldstein, ed., Hemisphere,
Washington, DC, pp. 559615.
8 Reiss, L. P., and Hanratty, T. J., 1963, An Experimental Study of the Unsteady Nature of the Viscous Sub-Layer, AIChE J., 9, pp. 154160.
9 Mitchell, J. E., and Hanratty, T. J., 1966, A Study of Turbulence at a Wall
Using an Electrochemical Wall Shear Stress Meter, J. Fluid Mech., 26, pp.
199221.
10 Fortuna, G., and Hanratty, T. J., 1971, Frequency Response of the Boundary
Layer on the Wall Transfer Probes, Int. J. Heat Mass Transfer, 14, pp. 1499
1507.
11 Deslouis, C., Gil, O., and Tribollet, B., 1990, Frequency Response of Electrochemical Sensors to Hydrodynamic Fluctuations, J. Fluid Mech., 215, pp.
85100.
12 Sobolik, V., Wein, O., and Cermac, J., 1987, Simultaneous Measurement of
the Film Thickness and Wall Shear Stress in Wavy Flow of Non-Newtonian
Liquids, Collect. Czech. Chem. Commun., 52, pp. 913928.
13 Labraga, L., Bourabaa, N., and Berkah, T., 2002, Wall Shear Stress From a
Rotating Cylinder in Cross Flow Using the Electrochemical Technique, Exp.
Fluids, 33, pp. 488496.
14 Zhuoxiong, M., and Hanratty, T. J., 1991, Analysis of Wall Shear Stress
Probes in Large Amplitude Unsteady Flows, Int. J. Heat Mass Transfer, 34,
pp. 281290.
15 Rehimi, F., Aloui, F., Ben Nasrallah, S., Doubliez, L., and Legrand, J., 2006,
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
16
17
18
19
20
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
M. A. El-Gebeily
x=
Introduction
The bipolar cylindrical coordinate system , , z is a threedimensional orthogonal coordinate system that results from projecting the two-dimensional bipolar coordinate system , , in
Contributed by the Fluids Engineering Division of ASME for publication in the
JOURNAL OF FLUIDS ENGINEERING. Manuscript received April 2, 2008: final manuscript
received March 11, 2009; published online April 16, 2009. Review conducted by
Dimitris Drikakis.
y=
a sin
,
cosh cos
and
z=z
h =
a
,
cosh cos
and
hz = 1 2
a sinh
,
cosh cos
a
,
cosh cos
h h zw =
h
h
+
=0
h h z
hhz
aUycosh cos 1
cosh cos 2
and
cosh cos 2
5
or simply,
aUy sinh
cosh cos
as
, 0,0
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
x=0
x = x2
x = x1
Ux
q=0
q=0
q=p
x=0
Uy
, =
cos n sinh n 1
n=1
, =
cn cos n sinh n 1 +
n=1
sinh n1
n=1
10
Similarly, at 1 we set
=
=1
nd
sin n sinh n1 2 =
n=1
, =
cos n sinh n 2
n=1
=
=2
nc
sin n sinh n2 1 =
n=1
and
, =
n=1
sinh n2
n=1
12
Using the orthogonality properties of trigonometric functions, the
coefficients cn and dn can be determined as
cn =
9
The fact that does not vanish as , 0 , 0 can be dealt with
dn cos n sinh n 2 +
aUy sinh 2
n sinh n2 1
n2
2aUy e
sinh n2 1
sin sin n
d
cosh 2 cos 2
13
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
0.5
0.25
0
-0.25
-0.5
-0.75
-1
0
0.5
1.5
2.5
aUy sinh 1
n sinh n1 2
dn =
sin sin n
d
cosh 1 cos 2
+n1
= +
2aUy e
sinh n1 2
14
cos ne
n=1
n=1
n2
+n1
sinh n 2
15
Typical streamline patterns are shown in Fig. 3 for both the transformed and actual physical domains. The case shown in Fig. 3 is
for r1 = 1, r2 = 2, and H = 4 where H is the center-to-center distance. The streamlines are orthogonal to the boundaries = 0 and
= . The two lines originally the cylinders = 1 and = 2 are
streamlines themselves. The point , 0 , 0 works like a
black hole where all streamlines are sucked into.
Once the flow field has been determined, the distribution of the
pressure P along the surfaces of the cylinders can be obtained
from the equation of conservation of momentum. The
-component of the momentum equation can be written as
1
U2y
w
d
17
16
1
U2y
d = 1 cos 2
19
n=N+1
w
P
= w
U2y
18
sinh n 1 + e
sinh n2 1
P P0
sinh n2 + 1
sinh n2 1
P , =
sinh n2
1 e22
20
cosh cos 2
and
cosh cos 2
21
or simply,
MAY 2009, Vol. 131 / 054501-3
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Fig. 4 The pressure distribution around the right cylinder for the case r1 = 1, r2 = 1
aUx sin
cosh cos
as
, 0,0
22
fn =
= +
aUx sin
f n sin n sinh n 1
cosh cos n=1
gn sin n sinh n 2
23
gn =
n=1
2aUxe
sinh n2 1
n2
x , = +
aUx
n sinh n2 1
aUx
n sinh n1 2
24
+n1
= +
2aUxe
sinh n1 2
25
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
1
0.8
0.6
0.4
0.2
0
-0.2
0.5
1.5
2.5
sin
x ,
= +
2aUx
2cosh cos
n=1
cosh cos 2
cosh cos 2
aUo cos sinh sin aUo sin cosh cos 1
cosh cos 2
cosh cos 2
27
Or simply,
cosh cos
cosh cos
as
, 0,0
28
Fig. 7 The pressure distribution around the right cylinder for the case r2 = 1, H = 6
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
0.5
-0.5
-1
-1.5
0
, = y , + x ,
29
Acknowledgment
The authors would like to thank King Fahd University of Petroleum & Minerals KFUPM for supporting this research under
Grant No. MS/Inviscid/375. R.S.A. is grateful to Professor M.
Amara and the Applied Mathematics Laboratory at the Pau University in France for such a pleasant visit that made this research
possible.
References
1 Alassar, R. S., Badr, H. M., and Allayla, R., 2000, Viscous Flow Over a
Sphere With Fluctuations in the Free-Stream Velocity, Comput. Mech.,
265, pp. 409418.
2 Badr, H. M., 1994, Oscillating Inviscid Flow Over Elliptic Cylinders With
Flat Plates and Circular Cylinders as Special Cases, Ocean Eng., 211, pp.
105113.
3 Alassar, R. S., and Badr, H. M., 1999, Oscillating Flow Over Oblate Spheroids, Acta Mech., 13734, pp. 237254.
4 Alassar, R. S., and Badr, H. M., 1999, Oscillating Viscous Flow Over Prolate
Spheroids, Trans. Can. Soc. Mech. Eng., 231A, pp. 8393.
Downloaded 03 Jun 2010 to 171.66.16.158. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm