You are on page 1of 11

Nuclear Engineering and Design 313 (2017) 214224

Contents lists available at ScienceDirect

Nuclear Engineering and Design


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

Vibration response of a pipe subjected to two-phase flow: Analytical


formulations and experiments
L. Enrique Ortiz-Vidal a,, Njuki W. Mureithi b, Oscar M.H. Rodriguez a
a
Department of Mechanical Engineering, Sao Carlos School of Engineering, University of Sao Paulo (USP), Av., Trabalhador So-carlense, 400, 13566-970 So Carlos, SP, Brazil
b
Department of Mechanical Engineering, Polytechnique Montreal, Dpartement de Gniemcanique 2900, H3T 1J7 Montreal, QC, Canada

h i g h l i g h t s

 Analytical formulations for two-phase flow-induced vibration (2-FIV) are presented.


 Standard deviation of acceleration pipe response is a function of the square of shear velocity.
 Peak frequency is correlated to hydrodynamic mass and consequently to void fraction.
 Dynamic pipe response increases with increasing mixture velocity and void fraction.
 Hydrodynamic mass in 2-FIV in horizontal pipe is proportional to mixture density.

a r t i c l e i n f o a b s t r a c t

Article history: This paper treats the two-phase flow-induced vibration in pipes. A broad range of two-phase flow con-
Received 18 July 2016 ditions, including bubbly, dispersed and slug flow, were tested in a clamped-clamped straight horizontal
Received in revised form 14 December 2016 pipe. The vibration response of both transversal directions for two span lengths was measured. From
Accepted 16 December 2016
experimental results, an in-depth discussion on the nature of the flow excitation and flow-parameters
influence is presented. The hydrodynamic mass parameter is also studied. Experimental results suggest
that it is proportional to mixture density. On the other hand, two analytical formulations were developed
Keywords:
and tested against experimental results. One formulation predicts the quadratic trend between standard
Pipe flow
Two-phase flow
deviation of acceleration and shear velocity found in experiments. The other formulation indicates that
Gas-liquid flow the peak-frequency of vibration response depends strongly on void fraction. It provides accurate predic-
Flow-induce vibration tions of peak-frequency, predicting 97.6% of the data within 10% error bands.
2016 Elsevier B.V. All rights reserved.

1. Introduction full understanding of the two-phase flow mechanisms, which still


is an open topic.
Flow-induced vibration is a common phenomenon in industry. On account of the emphasis of research on nuclear industry,
The flow generates dynamic fluid forces which excites the struc- where both axial and cross flows are predominant, the available
ture inducing mechanical vibrations. The phenomenon has been information on flow-induced vibration subject to two-phase pipe
studied extensively over the last four decades, mainly for nuclear flow is scanty (Chen, 1991; Ortiz-Vidal and Rodriguez, 2011;
industry applications (e.g. Fujita, 1990; Padoussis, 2006; Pettigrew et al., 1998). Experimental studies noticed influence of
Pettigrew et al., 1998; Weaver et al., 2000). To date, the phe- two-phase flow parameters such as, mixture velocity, void fraction
nomenon of structural vibration induced by single-phase flow and flow-pattern on the structural response of the pipe system. For
(FIV) is reasonably well understood. Researchers have aimed to example, vibration response increases with increasing mixture
reach that same level of understanding for vibrations due to two- velocity and homogeneous void fraction. It also depends on flow-
phase flow. This is a challenge considering that it depends on the pattern (Geng et al., 2012; Ortiz-Vidal et al., 2013; Zhang and Xu,
2010). The characteristics of bubbly flows can change due to the
level of flow-induced vibration and as a function of liquid superfi-
cial velocity; however, these changes are not big enough to alter
Corresponding author.
the flow pattern (Hibiki and Ishii, 1998). Despite the mentioned
E-mail addresses: leortiz@sc.usp.br, enrique.ovl@outlook.com (L.E. Ortiz-Vidal),
njuki.mureithi@polymtl.ca (N.W. Mureithi), oscarmhr@sc.usp.br (O.M.H. Rodri- findings, a comprehensive and deeper discussion on the effect of
guez). two-phase flow parameters on vibration response is in order. The

http://dx.doi.org/10.1016/j.nucengdes.2016.12.020
0029-5493/ 2016 Elsevier B.V. All rights reserved.
L.E. Ortiz-Vidal et al. / Nuclear Engineering and Design 313 (2017) 214224 215

Nomenclature

A cross-sectional inside pipe area, [m2] fT, fs, ftp total, structural and two-phase damping components
C damping coefficient, [Ns/m] [%]
d, D internal and external diameters, [m] l viscosity [Pas]
EI rigidity, [Nm2] q density [kg/m3]
J, J (mixture) velocity, shear velocity, [m/s] Tk,w wall shear stress of k-phase
Ji joint acceptance [adm] swall, slam, sturb, wall, laminar and turbulent shear stress [N]
Lspan length between clamps, [m] xi, xpeak, xair, natural (for ith mode), resonance-peak and in-air
m linear mass density, [kg/m] frequencies [Hz]
P pressure, [Pa] /(x), w(x) mode shapes of the pipe and the pressure field
p0 pressure fluctuation, [Pa]
Q volumetric flow rate, [m3/s] Subscripts
y(x,t) transversal displacement of the pipe [m] acc acceleration
STD, PSD standard deviation and power spectral density G, L gas (air), liquid (water)
r radius [m] h hydrodynamic
V velocity [m/s] H homogeneous
v velocity fluctuation [m/s] k k-phase
x quality [adm] Lspan length between clamps [m]
a void fraction [adm] M mixture
b homogeneous void fraction [adm] p pipe

nature of flow excitation and the effect of flow patterns, specially


slug-flow, and flow-pattern transition on vibration response
should be analyzed.
Dynamic response of a pipe subjected to two-phase flow can be
represented by damping, hydrodynamic mass and excitation
mechanisms (Fujita, 1990; Pettigrew and Taylor, 1994). These
dynamics parameters are also related to two-phase flow parame-
ters. Significant research efforts on the study of damping have been
spent in recent years (Bguin et al., 2009; Charreton et al., 2014;
Gravelle et al., 2007). Results indicate that both structural and
two-phase damping components are relevant in structural vibra-
tion induced by two-phase flow (2-FIV) in pipes. Furthermore, Fig. 1. Scheme of a pipe subjected to internal flow.
the latter has a viscous nature (velocity dependent) and is highly
dependent on void fraction. Hydrodynamic mass in two-phase pipe
flow, on the other hand, has not received much attention. No report vibration-excitation mechanisms, and consequently the vibration
on any experimental study has been found in the literature. In the response y(x,t). The turbulence and two-phase flow excitation
case of excitation mechanisms, both turbulence and two-phase mechanisms are the most relevant for both liquid and gas-liquid
flow are the most important mechanisms in two-phase pipe flow pipe flow (Pettigrew et al., 1998). Turbulence is associated to the
(Pettigrew and Taylor, 1994; Riverin and Pettigrew, 2007). An ana- presence of eddies of many sizes who decay to smaller ones, dissi-
lytical deduction has been reported in the literature to describe a pating energy. The eddies in the region near the pipe wall generate
relationship between vibration response and flow rate based on random pressure fluctuations that forces it to vibrate (Blevins,
the mechanism of turbulence (Evans et al., 2004). No other formu- 2001). In the case of two-phase flow, in addition to eddies, the
lation on vibration response including excitation mechanism has instant reconfiguration of the phases induces perturbations on
been found in the open literature. the flow. Depending of their nature, these perturbations can be
In this paper, flow-induced vibration subject to two-phase pipe another source of turbulence or can promote the emergence of a
flow is investigated. Experiments in a broad range of two-phase two-phase flow excitation mechanism. Thus, the relationship
flow conditions, including bubbly, slug and dispersed flow, in between the structural response and flow perturbations is evident.
two span lengths of a clamped-clamped straight horizontal pipe
are carried out. We present an in-depth discussion on 2-FIV vibra-
2.1. Relation between standard deviation of acceleration and shear
tion response from these experiments; e.g. the effects of flow-
velocity
pattern and hydrodynamic mass are presented. In addition, two
analytical formulations are developed and tested. One relates the
The development of an analytical formulation of flow-induced
vibration response to shear velocity based on the mechanical
vibration subjected to two-phase flow (2-FIV) in pipes has to do
energy equation of the two-phase flow and random vibration con-
with the establishment of a relationship between flow and struc-
cepts. The other formulation indicates the relationship between
ture parameters. To date, exact mathematical deductions are not
peak-frequency and void fraction.
available, mainly because some topics, e.g. turbulence, two-phase
flow and their interaction, are still in development. Here we ana-
2. Analytical approach lytically present a relationship between shear velocity and the
standard deviation of acceleration of the pipe. This is done in
Fig. 1 represents a pipe subject to internal adiabatic two-phase two steps. First, the mechanical energy equation of the flow is used
flow, where the flow characteristics determine the governing to relate pressure to shear velocity. Next, pressure and standard
216 L.E. Ortiz-Vidal et al. / Nuclear Engineering and Design 313 (2017) 214224

deviation of acceleration are correlated for the pipe, based on a ity perturbation, namely the Reynolds tensor. Therefore, there is a
random vibration approach. In the following analysis, we start clear relationship between two-phase wall shear stress and pres-
from fundamental equations for two-phase flow and structural sure perturbations that can be represented by
dynamics in order to establish relationships between flow and
f sw ; p0 0 4
structural parameters. Physical interpretations from previous
theoretical and experimental results and well-known standard
STDp0 c1 sw 5
mathematical methods are used.
The two-phase pipe flow of Fig. 1 can be modeled by the one- where c1 is a constant.
dimensional two-fluid model. The momentum equation of The perturbation component of pressure field in Fig. 1, i.e. p0 (x,t),
k-phase is represented by (Ishii and Hibiki (2006) and Rosa (2012)) acts as a distributed transverse load on the pipe system. It can be
assumed as stationary, zero mean, ergodic random process. Then,
@ @
hak iqk hV k i C Vk hak iqk hV k i2 the dynamic equation of motion can be expressed by
@t @x
@Pk Sk @4 @ @2
hak i  ak;w T k;w hV ki ihCk i hMki i 1 EI yx; t C yx; t qA 2 yx; t p0 x; t 6
@x A @x 4 @t @t
where the subscripts k, i and w refer to k-phase, interfacial and pipe where E is the elastic modulus, I is the second moment of area of the
wall. Angle brackets represent flow properties averaged over the tube section and qA is the linear mass density of the system, includ-
crosssectional area. a, q, P and V represent respectively void frac- ing added mass. The product EI is the flexural rigidity. C represents
tion, density, pressure and actual velocity. CVk is the distribution the damping coefficient. In presence of a linear system, the
parameter and computes the effect of the void and momentum- transversal displacement y(x,t), in terms of normal modal expan-
flux profiles on the cross-sectional area. A and S denote the cross- sions, is defined as,
sectional inside pipe area and wall wetted perimeter, respectively.
X
N
T represents the shear stress. The last two terms on the rhs of yx; t /i xyi t 7
Eq. (1) are related to interfacial forces and are associated with i1
momentum transfer due to mass transfer at the interface and the
where the summation is truncated after all important modes. /i(x)
stress acting at the interface, respectively. In the case of single-
and yi(t) represent the modal shape and modal coordinate, for the
phase flow, these two terms disappear and ak becomes equal to one.
ith mode, respectively. The former parameter is related to natural
The mechanical energy equation can be used to identify the
frequencies of the system. Blevins (1979a,b) provides a vade mecum
dependence between flow parameters. This energy equation is
for natural frequencies and modal shapes for different structural
obtained by multiplying the momentum equations (Eq. (1)) by
components. Substituting Eq. (7) in Eq. (6), multiplying by /j(x)
the respective average velocity, and adding the equations of both
and integrating over the length, the system equation becomes
phases. For most practical applications, the static pressure of the
RL
two phases can be considered equal (Pk = P) (Ishii and Hibiki, 1 2fi /i xp0 x; tdx
2006). Thus, we have the mechanical energy equation for two- t
y y_ i t yi t 0
R 8
xi 2 i xi qAx2i 0L /2i xdx
phase flow,
" # where xi and fi represent the natural frequency and viscous damp-
X2
@ hV k i2 X 2
@ hV k i2
hak iqk C Vk hak iqk hV k i hak ihV k iP ing factor, for ith mode, respectively. Eq. (8) assumes uncoupled sin-
k1
@t 2 k1
@x 2 gle dominant modes with a small damping ratio. In a similar
X
2 manner to the structural displacement, it is reasonable to separate
hV k ihF kM i 2 the pressure, p0 (x,t), into functions of space and time. In the follow-
k1 ing, we consider that pressure load contain a broad range of fre-
where FkM represents the source term. Considering small perturba- quencies and the majority of the pipe response due to the load is
tions in velocity and pressure, vk and p, upon an uniform steady contained at frequencies near the system natural frequencies. These
flow, Vk, P, respectively, and retaining only the first-order terms assumptions are coherent with several experimental results (e.g.
for simplicity (this does not detract generality), the mechanical Chung, 1985; Durant and Robert, 1998; Ortiz-Vidal et al., 2013).
energy equation becomes, Then, for each mode, pressure load can be expressed as
" ! # p0i x; t wi xp0i t 9
X
2
@ X2
@ 3V 2
ak qk V k v 0k ak C Vk qk k P v 0 0
V kp
where wi(x) is the shape of the distribution of the pressure load
k1
@t k1
@x 2 k

along the pipe, of the turbulence pressure load component in the


X
2
hV k v 0k ihF kM i 3 frequency band near the natural frequency of mode i. pi(t) repre-
k1 sents the load acting over the range of frequencies in the neighbor-
hood of the ith mode natural frequency. The last definitions are in
where the angle brackets are dropped for simplicity. For the present agreement with a previous approach by Blevins (2001). Substituting
analysis, void fraction perturbation is neglected. This assumption is Eq. (9) in Eq. (8),
reasonable for a stable flow pattern from the hydrodynamic point of
1 2fi
view, i.e., any void fraction perturbation does not promote transi- i t
y y_ i t yi t J i p0i t 10
tion to another flow pattern (Rodriguez and Bannwart, 2008; x2i xi
Rodriguez and Castro, 2014; Wallis, 1969). The phases are also
where Ji is the joint acceptance,
assumed incompressible. Eq. (3) is similar to the mechanical energy
RL
equation obtained by Laufer (1953) for turbulent single-phase pipe /i x wi xdx
flow. Eq. (3) establishes the interdependence between pressure and Ji 0
R 11
qAx2i 0L /2i xdx
velocity perturbations. For a number of systems of practical inter-
est, particularly for systems without mass transfer and phase Blevins (1989) studies different analytical approximations for Ji.
change, the source term FkM consists of dissipative surface forces, In general, the spatial distribution of pressure field, wi(x), is
specifically wall shear stress. In turn, FkM is also a function of veloc- unknown and experimental data on the pressure field is needed
L.E. Ortiz-Vidal et al. / Nuclear Engineering and Design 313 (2017) 214224 217

(Sankaran and Jancauskas, 1992). The joint acceptance equals unity 2.2. Relation between pipe peak frequency and void fraction
is commonly used for simplicity and corresponds to the case where
the pressure-field and the pipe-mode shapes are identical. Ji = 1 Pressure fluctuations, due to turbulence, excite the pipe system
produces a conservative estimate (Cunningham and White, 2004; in a wide range of frequencies corresponding to their infinite
Pettigrew and Taylor, 1994). Taking the Fourier transform of both degrees of freedom (modes) (Blevins, 1979a,b; de Silveira Neto,
sides of Eq. (10), 2002). Therefore, the structural response in frequency domain
shows a resonance phenomenon for each mode, i.e. the response
Y i x J i Hi xP0i x 12 is amplified in the neighborhood of the natural frequencies.
According to flow-induced vibration studies, the peak frequency
where Hi(x) is the transfer function of the system. It is expressed by value for each natural frequency is affected by different parameters
1 (Carlucci, 1980; Chung and Chen, 1984; Yi-min et al., 2010), such
H i x   13 as flow velocity and fluid density. In the case of two-phase flow,
x2
x2i
2ifi xx 1 fluid density depends on void fraction. Thus, a relationship
i
between peak frequency and void fraction (a) is reasonable. The
x is the frequency of the excitation. Eq. (12) can be expressed in following expressions can be used for this purpose,
term of the Power Spectral Density of the acceleration response s  1=2 q s2
EI mp cJ mh J
(PSDacc) by multiplying each term by its complex conjugate and xpeaki ki Lspan 2
  1  f2T  1 
using the PSD definition, mp L4span mp mh |{z} EIk i
2

|{z} |{z} III |{z}


II IV
I
PSDacci x J 2i x4 jHi xj2 PSDp0 i x 14
21
PSDp0 i represents the power spectral density of the perturbation
component of the pressure. The fact that turbulence acts in a broad mh cm qL A1  a 22
frequency spectrum suggests that the pipe is excited in a similar
where xpeaki is the peak frequency of the ith mode. The terms I on
range of frequencies. Moreover, experimental results show PSDp-i
the rhs of Eq. (21) is related to theoretical natural frequency, being
is nearly uniform over a wide-band of the frequencies (Durant
that the factor kiLspan is the roots of frequency equation and depends
and Robert, 1998). Considering a pressure-fluctuation spectrum
on boundary conditions of the pipe (Blevins, 1979a,b). Lspan and mp
with constant magnitude PSDp in the range x1 6 x 6 x2, standard
are the length and linear mass density of the pipe, respectively. The
deviation of acceleration (STDacc) for ith mode is obtained by inte-
terms II and III take into account the effect of hydrodynamic mass
grating Eq. (14) over this range,
(mh) (Chung and Chen, 1984) and damping (f), respectively. The
Z x2 Z x2 term IV accounts for the influence of velocity, where the parameter
STD2acci PSDacci x dx J 2i PSDp0 x4 jHi xj2 dx 15 cJ is 0.55 for the first mode and 1 for the others (disregarding the
x1 x1
Coriolis effect) (Yi-min et al., 2010). In Eq. (22), cm is the hydrody-
By definition, the PSD of the pressure fluctuations is related to namic mass coefficient and when it is equal to 1 the hydrodynamic
its standard deviation by means of (Yang, 1986), mass is proportional to mixture density.
The formulation presented in Eqs. (21) and (22) is relevant for
STD2p0 2x2  x1 PSDp0 16 two-phase pipe flow and complements the formulation of the Sec-
tion 2.1 (Eq. (20)). In single-phase flow, for constant geometry and
Combining Eqs. (16) and (15),
known roughness, wall shear stress is exclusively a function of flow
STDacci ci STP p0 17 velocity (White, 1998). Consequently, Eq. (20) is enough to estab-
lish a relationship between structural response and flow. By con-
 Z 1=2 trast, a unique relationship is insufficient in two-phase flow
x2
1 because of wall shear stress is function of other parameters, such
ci J i x4 jHi xj2 18
2x2  x1 x1 as void fraction and mixture velocity (Ortiz-Vidal et al., 2014).
Thus, Eqs. (21) and (22) are necessary. Both formulations are com-
The overall acceleration response is obtained by summing of
plementary and establish a relationship between structural
squares of the contributions of the individual modes (Au-Yang,
dynamic and two-phase void fraction and velocity parameters.
1975; Blevins, 1989). Then, STD of acceleration response at refer-
ence position x1 can be expressed by
3. Experimental work
!1=2
X
N
STDacc STDp0 f/x1 i  ci g2 19 3.1. Two-phase flow facilities
i1

Eq. (19) establishes that STD of acceleration is proportional to The experimental work was conducted at the Fluid-Structure
STD of pressure fluctuations. Furthermore, from Eq. (5), in turn, Interaction Laboratory of the BWC/AECL/NSERC Industrial Research
the latter is proportional to wall shear stress. Introducing shear Chair, Polytechnique Montreal. Fig. 2 shows the test loop used to
p
velocity (J  sw =qM ) as a more convenient parameter to express investigate flow-induced vibration subjected by gas-liquid two-
fluctuations in turbulence (e.g. Metzger and Klewicki, 2001) and phase pipe flow. The main components and measurement sensors
combining Eqs. (19) and (5), STDacc can be represent by are designated by letters and numbers and listed in Tables 1 and 2,
respectively. The experimental facility was especially designed and
STDacc c2 qM J 2 20 constructed according to Ortiz-Vidal (2012). A picture of the exper-
imental rig used in this study is shown in Fig. 3.
where c2 is a constant. Eq. (20) is the proposed analytical formula- Compressed air is supplied into the system and its pressure con-
tion based on turbulence that establishes a quadratic relationship trolled by the regulator 1, which maintained a steady pressure.
between standard deviation of acceleration and shear velocity. qM Two air flowmeters (2 and 3 in Fig. 2) and the valve A are used
represents the mixture density. to measure and regulate the flow rate, respectively. From an inde-
218 L.E. Ortiz-Vidal et al. / Nuclear Engineering and Design 313 (2017) 214224

slug frequency, is negligible. In this case, the adopted entrance


length ensures the development of intermittent flow patterns.

3.2. Test section

The test section (I, Fig. 2) consists of a clamped-clamped hori-


zontal PVC pipe with 20.4 and 26.7 mm of internal (d) and external
(D) diameters, respectively. The span is delimited by the distance
between heavy clamps (H and J, Fig. 2). In the present study, two
spans were tested, Lspan = 40d (=816 mm) and 75d (=1530 mm).
Impact test and circle-fitting method techniques were performed
to extract the natural frequency and damping of the system. Mea-
surements for the two transversal directions with both stagnant
water and air were done. Structural damping ratio ranged between
0.7% and 2.3%. Experimental natural frequencies are in agreement
with those predicted by continuous-system theory (Blevins, 1979a,
b), The percent error between measurements and theory is less
than 5%. For example, experimental and predicted first-mode nat-
ural frequencies of air-filled span of 75d are 16.29 and 16.49 Hz
(gravity direction), respectively. The following equation was used,
s
EI
xi ki Lspan 2
23
mP L4span
Fig. 2. Schematic representation of the test loop.
where EI = 39.47 Nm2. The factor kiLspan takes the values of 4.7300
and 10.9956 for the first and third modes, respectively. The linear
Table 1 mass density of the pipe, mp, is 0.336 kg/m.
Main components of the test loop. A triaxial accelerometer located at the midspan position is used
to monitor the vibration response of the test section. The x-axis is
Letter Component
parallel to the direction of the flow (axial direction). Both z and y
A Air regulation valve
axes are perpendicular to pipe axis and z-axis lies in the gravity
B Water reservoir
C Water centrifugal pump
direction. The sensitivity for x, y and z axes is 102.0 mV/g,
D Frequency shifter 102.7 mV/g e 99.7 mV/g, respectively.
E Water centrifugal pump
F Water regulation valve 3.3. Experimental matrix
G Mixer
H Heavy clamp
I Test section The experimental matrix is composed from 8 water single-
J Heavy clamp phase and 32 air-water two-phase flow conditions, for both test
sections of spans 40d and 75d. Flow velocities from 0.25 m/s to
7 m/s are used to water single-phase flow. In the case of air-
Table 2 water flow, two parameters are used: the mixture velocity,
Measurement sensors of the test loop. referred simply as velocity in this paper, J, and the homogeneous
Number Component Range Accuracy void fraction (or simply void fraction), b,
1 Pressure regulator 0100 psig QL QG
2 Air flowmeter 075 slpm 1% FS J JL JG 24
A
3 Air flowmeter 01000 slpm 1.5% FS
4 Pressure gage 0100 psig 3% FS
QG JG
5 Water flowmeter 10.18424.18 lpm 0.25% RD b 25
6 Pressure transducer 050 psig 0.05% FS Q L Q G JL JG
7 Pressure transducer 050 psig 0,05% FS
where QL, QG and A represent the liquid and gas flow rates and
pipes inner cross-sectional area, respectively. Every flow condition
pendent line, water at standard conditions is pumped from the is set for the midspan position of the test section, i.e. in-situ values.
reservoir B by the centrifugal pumps C and E. The water flow rate For this, the air flow rate is corrected from the pressure values
is measured and regulated by the flowmeter 5 and valve F, respec- downstream and upstream (6 and 7, Fig. 1), assuming a linear pres-
tively. Water and air are mixed in mixer G and the two-phase mix- sure variation along the test section. The mixture velocity and void
ture flowed through the pipe in several flow-patterns. Then, the fraction ranges from 0.5 m/s to 25 m/s and 10% to 95%, respectively.
fluids are separated by gravity in the atmospheric tank B. Tested experimental conditions include bubbly, dispersed and slug
An entrance length of 1760 mm (86d), between the mixer G flows as showed in Fig. 4. More details on experimental work are
and the pressure tap 6 (vide Fig. 2) is considered. This length given in Ortiz-Vidal et al. (2014).
ensures a fully developed flow for all tested single-phase flow con-
ditions (White, 1998). The adopted entry length is also enough for 3.4. Data acquisition and processing
dispersed two-phase flow conditions, provided that it can be con-
sidered as a pseudofluid with homogeneous properties (Wallis, A program developed in LabVIEWTM, via PXI National Instru-
1969). In the case of intermittent flows, elongated bubbly and slug, mentsTM acquisition boards, was used for acquiring and processing
a minimum length of 60d is needed (van Hout et al., 2003). From the signals. The sampling rate of 5000 samples/s was chosen to
this length value, the coalescence rate of liquids slugs, related to ensure the accuracy of the sensors and avoid the aliasing effect.
L.E. Ortiz-Vidal et al. / Nuclear Engineering and Design 313 (2017) 214224 219

Fig. 3. Experimental rig used in this study.

Fig. 4. Air-water experimental conditions on Mandhane et al. flow map (Mandhane et al., 1974). The solid grey, white and black symbols represent the bubbly, dispersed and
slug flow patterns, according to experimental observations (Ortiz-Vidal et al., 2014), respectively. The void fraction classification is based on the nominal value.

For each flow condition, 100-s time acceleration data were


acquired. The data were low-pass filtered at 750 Hz and windowed
in samples of four seconds with 50% overlap. Power Spectral Den-
sity (PSD) was computed for each sample and then averaged.
Finally, standard deviation of acceleration (STDacc) was calculated
from the averaged Power Spectral Density (simply PSDacc) using
Parsevals theorem (Allen and Mills, 2004).
Shear velocity was directly calculated from experimental
pressure-drop data collected for water single-phase and air-
water two-phase pipe flow, by means of Eq. (26) (White, 1998)
(pressure drop data can be found in Ortiz-Vidal et al. (2014)).
r s
sw d dP
J  26
qM 4qM dx

4. Results and discussion

4.1. Experimental observations

Fig. 5. Intensity time-frequency map for water single-phase flow excitation. The
4.1.1. Nature of flow excitation map shows the Zaxis response for the system span 75d at 7 m/s. Dark regions
Fig. 5 shows a representative intensity time-frequency map of indicate peak frequencies and correspond to natural frequencies for first, third and
the structure response under single-phase flow excitation. It is fifth modes.
220 L.E. Ortiz-Vidal et al. / Nuclear Engineering and Design 313 (2017) 214224

constructed from PSD of acceleration calculated based on samples The effect of other flow parameters is observed when analyzing
of 4 s time acceleration signal, where the grayscale intensity is nor- the pipe dynamic response due to two-phase flow. We found influ-
malized by minimum and maximum fixed arbitrary values. The ence of mixture velocity, homogeneous void fraction, flow pattern
map indicates two important matters. On one hand, as expected and two-phase damping, being that it is not possible to isolate the
(Blevins, 1979a,b; Chung, 1985), turbulence excites the structure effects as in the single-phase flow excitation case. Fig. 6b illustrates
in a wide range of frequencies, amplifying the response in the the findings. The influence of flow-pattern can be seen in the shape
neighborhood of the natural frequencies. Dark regions in Fig. 5 cor- of the response spectrum. The effect of slug-flow pattern is evident,
respond to the frequency values for first three modes. It is impor- see condition C24. Strong response for frequencies lower than
tant to note that the accelerometer was installed at midspan 10 Hz, corresponding to characteristic frequencies of liquid slugs,
position, then even modes were not captured. Turbulence, on the are observed. The response spectrum due to dispersed flow (C14)
other hand, generates a constant amplitude excitation for each is similar to single-phase flow excitation. Although bubbly flow
time interval, suggesting that the system is excited by the same (C9) is an intermittent flowpattern, the velocities are small and
force level. This fact is illustrated in Fig. 5 and was not reported the liquid-slug momentum transfer is not large enough to be cap-
before. Similar behavior was found in maps generated for two- tured by the accelerometer. Thus the response for bubbly flow has
phase flow in both spans 75d and 40d systems. These results are the same response as dispersed flow.
not presented here for the sake of brevity. The mixture velocity has also influence on peak frequency. For
same homogeneous void fraction, the value of peak frequency
decreases with increasing mixture velocity. This behavior is similar
4.1.2. Influence of flow parameters on acceleration-response spectra to single-phase flow excitation and noted in Fig. 6b when compar-
Fig. 6a illustrates the effect of flow velocity on dynamics ing conditions C9 and C14; both with b = 25%. On the effect of
response of the pipe system subjected to single-phase flow excita- velocity on the structural response level, two opposite behaviors
tion. The higher the flow velocity, the bigger the structural vibra- are found. For low homogeneous void fractions, in contrast to
tion. This may be explained as follows. As velocity is increased, single-phase flow, it is observed a slight variation on the level of
velocity fluctuations increase due to turbulence. Then pressure structure response, although C14 has a significantly higher veloc-
fluctuations are also more intense and consequently the structural ity, as seen in Fig. 6b. This effect is caused by higher values of
vibration. Moreover, the rate of increase of the structural response two-phase damping in dispersed flow, which reduces the level of
is more significant at higher velocities. This fact can be observed in the response (Bguin et al., 2009). On the other hand, for high
Fig. 6a, small variation of structural response is observed between homogeneous void fractions (b P 50), where slug flow occurs,
minimum (red line, J = 0.5 m/s) and medium (blue line, 3 m/s) higher variations of the structure response level are found.
velocities for a factor of 6 increase in flow velocity. By contrast, Homogeneous void fraction also acts on the pipe dynamics
when comparing the structural response of intermediate and max- response shifting the peak frequency to higher values when this
imum (black line, 7 m/s) velocities, where factor is lower (2.33), a flow parameter increases. This effect is noted in Fig. 6b when com-
larger variation is observed. This fact was also reported in the liter- paring conditions C9 or C14 (b = 25%) and C24 (b = 50%). The shift
ature (Jendrzejczyk and Chen, 1985) and can be related to turbu- of peak frequency is related to hydrodynamic mass and will be dis-
lence scales. cussed in the next section. From the discussion above, the pro-
Flow velocity also has influence on peak frequency. The value of posed peak-frequency analytical formulation appears to be
peak frequency decreases with increasing velocity, as shown consistent.
Fig. 6a. In this sense, flow velocity produces the same behavior as
a compressible load acting on the pipe. We showed recently 4.2. Hydrodynamic mass
(Ortiz-Vidal et al., 2013) that the frequency shift due to velocity
can be reduced or even eliminated through the normalization of Hydrodynamic mass (mh) calculations have been performed
frequency, where the dimensionless frequency is function of flow based on the Eq. (27) proposed by Carlucci (Carlucci, 1980). The
condition and critical buckling velocities. For general purposes, assumption of constant effective stiffness, implicitly in Eq. (27),
the velocity effect on frequency shift is negligible compared to is considered acceptable according to the discussion above, where
the velocity variation. a negligible influence of velocity was observed. The resonance fre-

Fig. 6. PSDacc for the first mode of 75d span system: (a) water single-phase flow excitation; (b) air-water two-phase flow excitation. C# corresponds to the flow conditions
shown in Fig. 4. The results are for Z-axis, which lies in the gravity direction.
L.E. Ortiz-Vidal et al. / Nuclear Engineering and Design 313 (2017) 214224 221

quencies of water single-phase flow or air-water two-phase flow sense, the equivalent mass of fluid vibrating with the structure
excitation, xpeak, were directly extracted from the PSDacc for each has similar behavior in both two-phase horizontal pipe flow and
flow condition. PSDacc and resonance frequencies for air (xair) were cross-flow excitations. On the other hand, the use of single-phase
obtained as described in the Experimental work section. fluid theory appears reasonable from a practical point of view. It
" # is more evident for the third mode, where lower scatter of the data
2
xair related to little influence of mass flux is observed.
mh mp 1 27
xpeak
Fig. 7 illustrates the variation of hydrodynamic mass normal- 4.3. Validation of the proposed analytical approach
ized by the corresponding single-phase flow value (mh,l) as a func-
tion of twophase flow parameters for the span-75d system. 4.3.1. Relationship between standard deviation of acceleration and
Hydrodynamic mass results as a function of void fraction, for the shear velocity
first and third modes and both transversal directions, are plotted The validation of proposed analytical formulations is presented
in Fig. 7a. It is observed that hydrodynamic mass decreases with in this section. First, the relationship between standard deviation
increasing void fraction, according to literature (e.g. Pettigrew of acceleration (STDacc) and shear velocity (J) is evaluated. Accord-
and Taylor, 1994). Furthermore, the results are in reasonable ing to the proposed formulation, there is a quadratic relation
agreement with single-phase flow theory, where hydrodynamic between both parameters. Thus, a quadratic fit, with its respective
mass is proportional to mixture density (cm = 1, Eq. (22)). Experi- coefficient of determination, R2, accompanies the results of STDacc
mental points are localized around the line established by theory. versus J. The influence of direction response and length span is also
Two more important matters should be noted. The difference assessed. For the former, results of both transversal directions (Y
between hydrodynamic mass in the Y and Z directions is not signif- and Z axis) of system span 75d are presented; for the latter, Z-
icant. On the other hand, the scatter of the data is larger in first- axis results for both systems, spans 40d and 75d, are compared.
mode results. This can be caused by the presence of two-phase Fig. 8 presents the results for low homogeneous void fractions,
flow phenomena, for example, liquid slugs with low frequencies including water single-phase flow (b = 0%). As expected, standard
affect the first-mode peak frequency response. Similar results are deviation of acceleration increases with increasing shear velocity.
found for the system span 40d. Reasonable good quadratic fits (R2 P 0.92) are observed, especially
Fig. 7b presents the hydrodynamic-mass ratio as a function of in two-phase flow (R2 P 0.98), indicating that the quadratic trend
mass flux. The set of air-water two-phase flow conditions of predicted by the proposed formulation is valid. Results suggest low
Fig. 4 are plotted. The hydrodynamic mass is practically indepen- influence of span length on STDacc. This behavior can be explained
dent of mass flux, except to b = 95%. Nevertheless, for constant by low shear velocities and, consequently, low levels of turbulence.
homogeneous void fraction, hydrodynamic mass slightly decreases The latter is not large enough to affect the pipe response level. It is
with increasing mass flux and then increases. Comparing Figs. 7b important to highlight the fact that the proposed formulation does
and 4, it is noted that the increment occurs when the dispersed- not establish the influence of span length.
flow-pattern boundary is reached. At this point, the hydrodynamic Other relevant matters should be observed in Fig. 8. Standard
coupling between pipe and flow is stronger, increasing the hydro- deviation of acceleration increases with increasing the homoge-
dynamic mass. The opposite happens with the homogeneous void neous void fraction, as previously reported by Geng et al. (2012)
fraction conditions of 95%; hydrodynamic-mass ratio decreases and Ortiz-Vidal et al. (2013). Also, STDacc for Y-axis is larger than
with increasing mass flux. This behavior can be attributed to inter- for Z-axis. This can be caused by the gravity field. The fluid mass
mittent flow, for which the hydrodynamic coupling is weak. For would be acting in Z-axes as a natural damper reducing the vibra-
75% of homogeneous void fraction, the presence of both opposite tion response. Concerning the shape of the fit, experimental points
effects described above can explain the almost constant value of follow a concave-up quadratic trend. An approach based on flow
hydrodynamic mass. pattern can be useful to explain this fact. For low homogeneous
The hydrodynamic mass results illustrated in Fig. 7 were not void fraction (10% and 25%), according to Ortiz-Vidal et al.
reported before. They are similar to results reported for cross- (2014), bubbles break gradually into small bubbles with increasing
flow, e.g. Olala et al. (2014) and Pettigrew et al. (1989). In this the mixture velocity. The turbulence increases progressively and

Fig. 7. Hydrodynamic mass ratio as a function of (a) void fraction and (b) mass flux. Void fraction is calculated using Armand-Massina correlation, a = (0.833 + 0.167x)b,
where x is quality. Room temperature value of 20 C is used. The results are for the system span 75d, where open and solids symbols correspond to first and third modes,
respectively. In (b), the homogeneous void fraction (b) classification is based on the nominal value.
222 L.E. Ortiz-Vidal et al. / Nuclear Engineering and Design 313 (2017) 214224

Fig. 8. Standard deviation of acceleration as a function of shear velocity for low homogeneous void fraction, including water single-phase flow (b = 0%). The systems span 75d
and 40d are represented by square and triangular dots, respectively. Open and solid dots corresponds to Y and Z axes, respectively. The dashed line represents the quadratic
trend according to formulation proposed. R2 is its coefficient of determination. Vertical solid lines indicate flow-pattern boundary transitions as shown in Fig. 4.

excites the pipe in the same way. The last explanation is phe- may be due to the presence of high shear velocities and an inter-
nomenologically compatible with experimental results by mittent flow pattern, which excites the pipe systems in different
Giraudeau et al. (2013). Measurements of flow-induced forces in manners.
an elbow show the same concave-up quadratic trend at bubbly- Fig. 9 also illustrates the effect of flow pattern and flow-pattern
to-dispersed flow transition. Thereby, the pipe system responds transition. On one hand, values of standard deviation of accelera-
to this specific flow-excitation with a concave-up trend. tion are higher for slug and dispersed flow patterns than for bubbly
Fig. 9 presents the results for high homogeneous void fractions. flow, which can be attributed to periodic changes in the systems
Similarly, (i) STDacc increases with increasing shear velocity, as mass (liquid-slug momentum exchange) and intense turbulence,
expected by theory, (ii) STDacc for Y-axis is slightly higher than respectively. This findings are in accordance with previous studies
for Z-axis and (iii) STDacc increases with increasing the homoge- (Geng et al., 2012; Ortiz-Vidal et al., 2013). On the other hand, a
neous void fraction. Good coefficients of determination concave-down trend is observed in the presence of flow-pattern
(R2 P 0.96) demonstrate that the quadratic trend predicted by transition. It can be related to the flow-induced forces. According
the proposed formulation is consistent. By contrast with low to Giraudeau et al. (2013), the two-phase flow excitation increases
homogeneous void fraction, some influence of span length on with increasing mixture velocity with a greater rate in slug flow
STDacc is observed, especially for b = 50% and 75%. This behavior than in dispersed flow. The two-phase damping can be also respon-

Fig. 9. Standard deviation of acceleration as a function of shear velocity for high homogeneous void fraction. The systems span 75d and 40d are represented by square and
triangular dots, respectively. Open and solid dots corresponds to Y and Z axes, respectively. The dashed line represents the quadratic trend according to formulation proposed.
R2 is its coefficient of determination. Vertical solid lines indicate flow-pattern boundary transitions as shown in Fig. 4.
L.E. Ortiz-Vidal et al. / Nuclear Engineering and Design 313 (2017) 214224 223

Fig. 11. Comparison of measured (xpeak-experimental) and predicted (xpeak-predicted)


Fig. 10. Peak-frequency deviation as a function of mixture velocity for several two- peak frequency for the third-mode of both systems, spans 75d and 40d. Open and
phase damping ratios. Eqs. (21) and (22) were used to calculated xpeak. xpeak-ref solid dots correspond to Y and Z axes, respectively. The homogeneous void fraction
corresponds to peak frequency with zero two-phase damping (ftp = 0%). The classification is based on the nominal value, according to Fig. 4. The dashed line
following values were used: fs = 1.5% and cm = 1. indicates 10% of percent error.

sible for a lower response in dispersed flow, according to Bguin


et al. (2009). This explains the trend for 50% and 75% homogeneous of both systems, spans 75d and 40d. Experimental peak frequencies
void fractions. In the case of 90%, the concave-down trend can be were extracted directly from the PSDacc for each flow condition. In
caused by the proximity of transition to annular flow pattern, the case of the predictions, and based on discussion above, the fol-
where the flow-induced forces start to decrease because of the lowing values were used: cm = 1, fs = 1.5% (measured mean value)
small variation of instantaneous homogeneous-void-fraction and ftp = 2.5% (arbitrary). As observed, the proposed formulation
(Giraudeau et al., 2013). predicts the experimental peak-frequencies with good accuracy,
97.6% of the data within 10% of point spreading. This fact demon-
strates that the relationship between peak frequency and void frac-
4.3.2. Relationship between peak frequency and void fraction
tion predicted by Eqs. (21) and (22) is consistent.
In the second part of this section, Eqs. (21) and (22), the pro-
posed formulation, which relates peak frequency (xpeak) to void
fraction (a), is tested. First, the matters of hydrodynamic mass 5. Conclusions
and damping should be discussed. As demonstrated in Section 4.2,
the use of single-phase fluid theory is reasonable to estimate Flow-induced vibration subject to two-phase pipe flow is the
hydrodynamic mass, especially for the third mode. This means that main goal of this paper. Analytical formulations that relate flow
this parameter (Eq. (22)) can be calculated using cm = 1. In the case to dynamic response parameters are presented and tested. Two
of damping, on the other hand, the literature reports that the total span lengths of a clamped-clamped straight horizontal pipe were
damping ratio (fT, Eq. (21)) for pipe flow can be considered as the subjected to a broad range of two-phase flow conditions, including
sum of both structural (fs) and two-phase (ftp) damping compo- bubbly, dispersed and slug flow patterns. The nature of the flow-
nents (Bguin et al., 2009; Gravelle et al., 2007). Furthermore, the excitation and flow-parameters influence and hydrodynamic mass
latter component ranging from 1 to 3.5% for a pipe diameter simi- were deduced from the vibration response. The following remarks
lar to the present study. Note that here only structural damping summarize the conclusions of the present study.
component was obtained from experiments.
We analyzed the effect of two-phase damping on peak fre- (1) Vibration response depends on flow and structural parame-
quency predicted by Eq. (21). Simulations with two-phase damp- ters such as mixture velocity, homogeneous void fraction
ing ratios of 0%, 1.5%, 2.5 and 3.5% were carried out for two- and flow pattern. The effect of two-phase damping and
phase flow conditions of Fig. 4. For each damping ratio, hydrody- hydrodynamic mass is also observed. Constant stiffness
namic mass was calculated by means of Eq. (22) and the appears to be a valid assumption.
Armand-Massina correlation (see Fig. 7). These results were used (2) Turbulent flow excites the structure in a wide range of fre-
to obtain the peak frequency (xpeak, Eq. (21)). Structural parame- quencies at a constant amplitude excitation level.
ters presented in the Experimental work section were used. (3) Standard deviation of acceleration (STDacc) is a function of
Fig. 10 presents peak-frequency deviations, taking as reference the square of shear velocity, corroborating the proposed ana-
the peak frequency with zero two-phase damping. As can be lytical formulation. Furthermore, STDacc increases with
observed, there is very small influence of two-phase damping on increasing homogeneous void fraction.
the proposed peak-frequency formulation. It is noteworthy that (4) Shear velocity appears to be an appropriate velocity scale to
this fact does not indicate the irrelevance of two-phase damping relate to the vibration response. It captures the flow phe-
on flow-induced vibration. Rather, other parameters have more nomena at the wall, where turbulence excitation mechanism
impact on the proposed peak-frequency formulation, such as happens.
hydrodynamic mass. (5) Peak frequency is strongly correlated to hydrodynamic mass.
Fig. 11 presents the performance of the proposed peak- The proposed peak-frequency formulation has been consis-
frequency formulation against measured data for the third-mode tent and accurate to predict it.
224 L.E. Ortiz-Vidal et al. / Nuclear Engineering and Design 313 (2017) 214224

(6) The use of single-phase fluid theory is reasonable to esti- Giraudeau, M., Mureithi, N.W., Pettigrew, M.J., 2013. Two-phase flow-induced
forces on piping in vertical upward flow: excitation mechanisms and
mate hydrodynamic mass, especially for the third mode.
correlation models. J. Pressure Vessel Technol. 135, 30907.
Therefore, a hydrodynamic mass coefficient cm = 1 can be Gravelle, A., Ross, A., Pettigrew, M.J., Mureithi, N.W., 2007. Damping of tubes due to
adopted for two-phase pipe flow. internal two-phase flow. J. Fluids Struct. 23, 447462.
(7) Hydrodynamic mass due to two-phase pipe flow excitation Hibiki, T., Ishii, M., 1998. Effect of flow-induced vibration on local flow parameters
of two-phase flow. Nucl. Eng. Des. 185, 113125.
has the same behavior as cross-flow excitation, for which Ishii, M., Hibiki, T., 2006. Thermo-Fluid Dynamics of Two-Phase Flow. Springer.
this dynamic parameter is practically independent of mass Jendrzejczyk, J.A.A., Chen, S.S.S., 1985. Experiments on tubes conveying fluid. Thin-
flux. Walled Struct. 3, 109134.
Laufer, J., 1953. The Structure of Turbulence in Fully Developed Pipe Flow (No.
NACA-TR-1174). Washington, DC, United States.
There are still some aspects demanding further investigation, Mandhane, J.M., Gregory, G.A., Aziz, K., 1974. A flow pattern map for gasliquid flow
such as the effect of the boundary and length span on the level in horizontal pipes. Int. J. Multiph. Flow 1, 537553.
Metzger, M.M., Klewicki, J.C., 2001. A comparative study of near-wall turbulence in
of pipe response. The latter is not established by the proposed high and low Reynolds number boundary layers. Phys. Fluids 13, 692.
STDacc formulation. The influence of two-phase flow pattern should Olala, S., Mureithi, N.W., Sawadogo, T., Pettigrew, M.J., 2014. Streamwise fluidelastic
be also quantified. forces in tube arrays subjected to two-phase flows. ASME 2014 Pressure Vessels
and Piping Conference. Volume 4: Fluid-Structure Interaction. ASME, p.
V004T04A014.
Acknowledgment Ortiz-Vidal, L.E., 2012. Design of an Experimental Facility for Two-Phase Pipe Flow
(Diseo de un banco experimental para la generacin de flujo bifsico agua-
aire). Pontifical Catholic University of Peru (PUCP).
The authors are grateful to Benedict Besner, Thierry Lafrance
Ortiz-Vidal, L.E., Mureithi, N., Rodriguez, O.M.H., 2014. Two-phase friction factor in
and Cedric Bguin for their support in the experimental facilities gas-liquid pipe flow. Therm. Eng. 13, 8188.
and discussions, and to Irma Consuelo Aguilar Avila for helping Ortiz-Vidal, L.E., Rodriguez, O.M.H., 2011. Flow-induced vibration due to gas-liquid
in data collect. L. Enrique Ortiz Vidal gratefully acknowledges the pipe flow: knowledge evolution. 21st Brazilian Congress of Mechanical
Engineering - COBEM2011. ABCM, Natal, RN.
support of the Petrobras/CNPq and of BWC/AECL/NSERC Chair of Ortiz-Vidal, L.E., Rodriguez, O.M.H., Mureithi, N.W., 2013. An Exploratory
Fluid-Structure Interaction, Polytechnique Montreal. Oscar M. H. Experimental Technique to Predict Two-Phase Flow Pattern From Vibration
Rodriguez is grateful to CNPq for the research grant. Response. ASME2013 Pressure Vessels and Piping Conference. Volume 4: Fluid-
Structure Interaction. ASME, Paris, p. V004T04A061.
Padoussis, M.P., 2006. Real-life experiences with flow-induced vibration. J. Fluids
References Struct. 22, 741755.
Pettigrew, M.J., Taylor, C.E., 1994. Two-phase flow-induced vibration: an overview.
Allen, R.L., Mills, D., 2004. Signal Analysis: Time, Frequency, Scale, and Structure. J. Pressure Vessel Technol., Trans. ASME 116, 233253.
Wiley-IEEE Press. Pettigrew, M.J., Taylor, C.E., Fisher, N.J., Yetisir, M., Smith, B.A.W., 1998. Flow-
Au-Yang, M.K., 1975. Response of reactor internals to fluctuating pressure forces. induced vibration: recent findings and open questions. Nucl. Eng. Des. 185,
Nucl. Eng. Des. 35, 361375. 249276.
Bguin, C., Anscutter, F., Ross, A., Pettigrew, M.J., Mureithi, N.W., 2009. Two-phase Pettigrew, M.J., Taylor, C.E., Kim, B.S., 1989. Vibration of tube bundles in two-phase
damping and interface surface area in tubes with vertical internal flow. J. Fluids cross-flow: part 1hydrodynamic mass and damping. J. Pressure Vessel
Struct. 25, 178204. Technol. 111, 466.
Blevins, R.D., 1979a. Formulas for Natural Frequency and Mode Shape. Van Riverin, J.-L., Pettigrew, M.J., 2007. Vibration excitation forces due to two-phase
Nostrand Reinhold, New York, NY, United States. flow in piping elements. J. Pressure Vessel Technol. 129, 713.
Blevins, R.D., 1979b. Flow-induced vibration in nuclear reactors: a review. Prog. Rodriguez, O.M.H., Bannwart, A.C., 2008. Stability analysis of core-annular flow and
Nucl. Energy 4, 2549. neutral stability wave number. AIChE J. 54, 2031.
Blevins, R.D., 1989. An approximate method for sonic fatigue analysis of plates and Rodriguez, O.M.H., Castro, M.S., 2014. Interfacial-tension-force model for the wavy-
shells. J. Sound Vib. 129, 5171. stratified liquidliquid flow pattern transition. Int. J. Multiph. Flow 58, 114
Blevins, R.D., 2001. Flow-Induced Vibration. Van Nostrand Reinhold, New York. 126.
Carlucci, L.N., 1980. Damping and hydrodynamic mass of a cylinder in simulated Rosa, E.S., 2012. Escoamento Multifsico Isotrmico. ARTMED EDITORA S.A, Porto
two-phase flow. J. Mech. Des. 102, 597602. Alegre.
Charreton, C., Bguin, C., Ross, A., Etienne, S., Pettigrew, M.J., 2014. Two-phase Sankaran, R., Jancauskas, E.D., 1992. A technique for the measurement of the cross
damping in vertical pipe flows: effect of void fraction, flow rate and external wind joint acceptance function. In: University of Tasmania (Ed.), 11th
excitation. ASME 2014 Pressure Vessels and Piping Conference. Volume 4: Australasian Fluid Mechanics Conference. Hobart, Australia, pp. 973976.
Fluid-Structure Interaction. ASME, p. V004T04A048. de Silveira Neto, A. , 2002. Fundamentos da Turbulncia nos Fluidos, in: Associaao
Chen, S.S., 1991. Flow-induced vibrations in two-phase flow. J. Pressure Vessel Brasileira de Cincias Mecnicas (ABCM) (Ed.), Coleao Cadernos de
Technol. 113, 234241. Turbulncia. ABCM, Rio de Janeiro, pp. 152.
Chung, H., 1985. Flow-induced vibrations of a circular cylinder due to turbulent van Hout, R., Shemer, L., Barnea, D., 2003. Evolution of hydrodynamic and statistical
internal flows. In: Offshore Mechanics and Arctic Engineering Symposium. parameters of gasliquid slug flow along inclined pipes. Chem. Eng. Sci. 58,
Dallas, TX, USA, pp. 398402. 115133.
Chung, H., Chen, S.S., 1984. Hydrodynamic mass (Book No. CONF-8406479). Wallis, G.B., 1969. One-Dimensional Two-Phase Flow. McGraw-Hill.
Cunningham, P.R., White, R.G., 2004. A review of analytical methods for aircraft Weaver, D.S., Ziada, S., Au-Yang, M.K., Chen, S.S., Padoussis, M.P., Pettigrew, M.J.,
structures subjected to high-intensity random acoustic loads. Proc. Inst. Mech. 2000. Flow-induced vibrations in power and process plant components
Eng., Part G: J. Aerosp. Eng. 218, 231242. progress and prospects. J. Pressure Vessel Technol. 122, 339348.
Durant, C., Robert, G., 1998. Vibro-acoustic response of a pipe excited by a turbulent White, F.M., 1998. Fluid Mechanics, McGraw-Hill series in mechanical engineering.
internal flow. Flow Turbul. Combust. 61, 5569. McGraw-Hill Higher Education.
Evans, R.P., Blotter, J.D., Stephens, A.G., 2004. Flow rate measurements using flow- Yang, C.Y., 1986. Random Vibration of Structures. Wiley-Interscience, New York.
induced pipe vibration. J. Fluids Eng. 126, 280. Yi-min, H., Yong-shou, L., Bao-hui, L., Yan-jiang, L., Zhu-feng, Y., 2010. Natural
Fujita, K., 1990. Flow-induced vibration and fluid-structure interaction in nuclear frequency analysis of fluid conveying pipeline with different boundary
power plant components. J. Wind Eng. Ind. Aerodyn. 33, 405418. conditions. Nucl. Eng. Des. 240, 461467.
Geng, Y., Ren, F., Hua, C., 2012. An auxiliary measuring technology of wet gas flow Zhang, M., Xu, J., 2010. Effect of internal bubbly flow on pipe vibrations. Sci. China
based on the vibration signals of the pipe. Flow Meas. Instrum. 27, 113119. Technol. Sci. 53, 423428.

You might also like