You are on page 1of 20

This article was downloaded by: [North Carolina State University]

On: 30 November 2014, At: 16:22


Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered
office: Mortimer House, 37-41 Mortimer Street, London W1T 3JH, UK

Journal of Hydraulic Research


Publication details, including instructions for authors and
subscription information:
http://www.tandfonline.com/loi/tjhr20

Numerical and experimental


investigation of transient pipe flow
a b c
Amgad S. Elansary , M. Hanif Chaudhry & Walter Silva
a
Irrigation and Hydraulics Dept., Faculty of Engineering , Cairo
University , Giza, Egypt
b
Dept. of Civil and Environmental Engineering , Washington
State University , Pullman, WA
c
Dept. of General Engineering , University of Puerto Rico ,
Mayguez, Puerto Rico
Published online: 13 Jan 2010.

To cite this article: Amgad S. Elansary , M. Hanif Chaudhry & Walter Silva (1994) Numerical and
experimental investigation of transient pipe flow, Journal of Hydraulic Research, 32:5, 689-706,
DOI: 10.1080/00221689409498709

To link to this article: http://dx.doi.org/10.1080/00221689409498709

PLEASE SCROLL DOWN FOR ARTICLE

Taylor & Francis makes every effort to ensure the accuracy of all the information (the
“Content”) contained in the publications on our platform. However, Taylor & Francis,
our agents, and our licensors make no representations or warranties whatsoever as to
the accuracy, completeness, or suitability for any purpose of the Content. Any opinions
and views expressed in this publication are the opinions and views of the authors,
and are not the views of or endorsed by Taylor & Francis. The accuracy of the Content
should not be relied upon and should be independently verified with primary sources
of information. Taylor and Francis shall not be liable for any losses, actions, claims,
proceedings, demands, costs, expenses, damages, and other liabilities whatsoever
or howsoever caused arising directly or indirectly in connection with, in relation to or
arising out of the use of the Content.

This article may be used for research, teaching, and private study purposes. Any
substantial or systematic reproduction, redistribution, reselling, loan, sub-licensing,
systematic supply, or distribution in any form to anyone is expressly forbidden. Terms
& Conditions of access and use can be found at http://www.tandfonline.com/page/
terms-and-conditions
Downloaded by [North Carolina State University] at 16:22 30 November 2014
Numerical and experimental investigation of transient
pipe flow
Etude numérique et expérimentale d'un écoulement
transitoire dans une conduite

AMGAD S. ELANSARY
Assistant Professor,
Irrigation and Hydraulics Dept., WALTER SILVA
Faculty of Engineering, Assistant Professor,
Cairo University, Dept. of General Engineering,
Downloaded by [North Carolina State University] at 16:22 30 November 2014

Giza, Egypt University of Puerto Rico,


Mayguez, Puerto Rico

M. HANIF CHAUDHRY
Professor,
Dept. of Civil and Environmental Engineering,
Washington State University,
Pullman WA

ABSTRACT
Two mathematical formulations for the computation of transient flow in piping systems are compared with
experimental data. The formulations are: a four-equations fluid structure interaction model (FSI) that includes
Poisson coupling, and a two-equations model for the fluid. Both models are solved numericaly using the
method of characteristics. A partial-closure of a valve located at an intermediate point in a pipeline is used to
create transient flow. The two-equations model computed the maximum pressure peak satisfactorily but the
FSI model gave an overall better simulation. An unsteady-friction model, added to the FSI model, did not
influence the final results significantly. The experimental procedures followed to obtain the valve characteris­
tics and the pressure history along the pipeline are explained in detail. Excellent numerical results at the valve
are obtained when experimental data is used to simulate the time-dependent boundary condition.

RÉSUMÉ
Deux formulations mathématiques pour Ie calcul d'un écoulement transioire dans des systèmes de conduites
sont comparées a des résultats expérimentaux. Les formulations sont un modèle d'intéraction fluide - structure
(FSI) a quatre éguations qui comprend Ie couplage de Poisson, et un modèle a deux equations pour Ie fluide.
Les deux modèles sont résolus numériquement par la methode des cractéristiques. Une fermeture partielle
d'une vanne située a un point intermediaire de la conduite est utilisée pour créer un transitoire. Le modèle a
deux equations a calculé de facon satisfaisante le pic de pression, mais le modèle FSI a donné une simulation
encore meilleure. Un modèle de frottement non permanent, ajouté au modèle FSI, n'a pas influence significa-
tivement les résultats finaux. Les procédures expérimentales suivies pour obtenir les cractéristiques de la
vanne de fermeture et revolution de la pression le long de la conduite sont présentées en détail. D'excellents
résultats numériques au niveau de la vanne sont obtenus lorsque les données expérimentales sont utilisées
comme condition a la limite en fonction du temps.

Revision received April 26, 1994. Open for discussion till April 30, 1995.

JOURNAL OF HYDRAULIC RESEARCH, VOL. 32, 1994, NO. 5 689


Introduction
The transient pipe flow can be computed by neglecting or including the dynamic response of the
pipeline. In the first case, the momentum and continuity equations for the fluid are solved simulta­
neously without considering the circumferential and axial motion of the pipe walls. Numerical
models for the solution of this two-equations formulation are available and have been extensively
verified for a wide range of applications (Chaudhry, 1987; Wylie and Streeter, 1993). When the
interactions between hydraulic and structure conditions are to be considered in the analysis, the
structural equations must be solved simultaneously with the fluid equations. Once the structural
effects are introduced the complexity of the model increases and the numerical solution becomes
more demanding. Wiggert et al. (1985, 1987) developed a one dimensional fluid-structure interac­
tion model that reduced the number of equations to four.
The fluid friction appears as a source of damping in the momentum equation. It is well-known that
Downloaded by [North Carolina State University] at 16:22 30 November 2014

the steady-state equation for friction do not simulate the shear stress accurately during transient
flow. Zielke (1968) developed an expression for unsteady laminar fluid friction. Trikha (1975) pro­
posed a simplification of Zielke's equation that have been used as an approximation for unsteady
friction computations in turbulent transient flow (Vardy, 1992; Budny et al., 1990).
Wiggert et al. (1985) and Budny et al. (1990) validated their fluid-structure interaction models by
using a sudden closure of a valve as the transient generator. Also, most of the new methodologies
for the computations of unsteady flow friction in laminar and turbulent flow have been verified
with experimental data for sudden valve closures (Zielke, 1968; Suo and Wylie, 1989; Vardy and
Hwang, 1991). In numerical simulations the valve was assumed to close completely in less than
one calculation time interval. The major attributes of a sudden valve closure as a boundary condi­
tion to generate waterhammer are the creation of a well-defined pattern of pressure heads with time
and the easy handling of the boundary conditions in the numerical computations. Possible limita­
tions of a boundary condition like this in the verification of certain numerical models is the fact that
the flow descelerates to zero velocity and reverses direction very quickly, not allowing a sustained
transient flow in the forward direction. Also, the boundary condition at the valve becomes inde­
pendent of time and the application of some numerical models could be restricted by changes at the
boundary. Finally, a sudden closure is not a common situation in real life applications because all
the flow regulating devices take some time to complete the movement.
The pressure oscillations generated by a partial valve-closure depend on the valve geometry, the
valve location in the system, the valve movement sequence and the percentage of closure. A partial
valve closure maintains flow in the forward direction and turbulent conditions for a longer period of
time, when compared to a total closure. If the valve is located at the downstream end of a pipe, with
its end open to the atmosphere or to a constant-head tank, any partial-closure movement will pro­
duce only one single pressure peak. If the valve is located in an intermediate position (hereafter
called intermediate valve) between two constant head-tanks, pressure oscillations can be generated
depending upon the percentage of closure and the distance from the valve to both tanks. These
oscillations will damp faster than those occuring from a total closure.
This paper presents experimental results obtained during transient conditions created by a control­
led partial-closure of an intermediate valve. This particular configuration provides a boundary that
maintains turbulent flow and can produce pressure oscillations for several cycles. At the same time,
no severe flow reverse is allowed because the valve remains partially open at all times. A motor­
ized butterfly valve with these characteristics was assembled and installed in an experimental
setup built in the R. L. Albrook Hydraulics Laboratory at Washington State University.

690 JOURNAL DE RECHERCHES HYDRAULIQUES, VOL. 32, 1994, NO. 5


The objective is to compare the prediction of pressure oscillations obtained using the two-equations
model, the four-equations model (FSI) with steady friction, and the FSI with the unsteady friction
term.

Numerical Simulations
Two formulations are used to solve the governing differential equations. Experimental information
at the valve is used to model this boundary condition. The treatment of the boundary condition is
described in the next section.
The first formulation is described in Chaudhry (1987). The equation of motion and continuity for
one-dimensional unsteady pipe flow are given by:
Downloaded by [North Carolina State University] at 16:22 30 November 2014

with

xw, = p,/uM/8 0)
where u = average flow velocity; p = pressure intensity; R = inside pipe radius; xws - steady state
shear stress at the wall; pf= fluid density; af= wave speed in fluid, and ƒ = the Darcy-Weisbach
friction factor.

By using the method of characteristics, the two partial differential equations, (Eqs. 1 and 2), are
transformed into the following ordinary differential equations:
dp du xws (4)

dx
which is valid along the positive characteristic line given by — = +af and,
dp du t,„ /5)
dt * ' dt ' R
dx
which is valid along the negative characteristic line given by — = -a f .
dt '
The second formulation is an axially-coupled model presented by Wiggert et al. (1985). Four par­
tial differential equations are solved for the state variables: fluid pressure, p , fluid axial velocity, v,
pipe axial velocity, ü, and the axial stress, ax. In the following analysis, an axisymmetric, linear
elastic pipe wall material, small deformation and a thin-wall pipe are assumed for the structural
equations. The fluid is described as being linear, homogeneous and isotropic and the pressure
always remains above vapor pressure. With these assumptions the governing equations becomes

at ox ax

dv dp KD _ (-i\

JOURNAL OF HYDRAULIC RESEARCH, VOL. 32, 1994, NO. 5 691


dax du KD - (8)
dx dt " A,

döx R dp „du n (9)


at e at dx
where
2RK.. 2, (10)
K* K/ +
7F(,-V)
Af and A, are the cross-sectional areas for the fluid and the pipe respectively; E is the Young modu­
Downloaded by [North Carolina State University] at 16:22 30 November 2014

lus of elastisity for the pipe wall; K is the bulk modulus of fluid; v is the Poisson ratio and e is the
pipe wall thickness.

The above four fluid-structure equations are coupled through the Poisson ratio, v, as given by
Equations 6 and 9. The method of characteristics is used to transform the linear, first-order, hyper­
bolic partial differential equations, (Eqs. 6-9), to ordinary differential equations. Following For-
sythe and Wasow (1960) Equations 6-9 can be solved to obtain the coupled wave speed for the
fluid and the pipe. The relation for the wave speed in fluid is

l
C) = -[M-,jM2-4a2fa2] (11)

and the axial wave speed in the pipe is

C) = l
-[M+,jM2-4afa2] (12)

where,

M = a~f+a~t + 2v K*R/ep,
2 K*
a, = —
1
Pf
2 E
=
°' P,
The four compatibility equations are given as follows:
1 vR - X 2t„ (13)
±C G C G ±C — + ■—(j/ ^ + C]G -
-dt
t f f^f-P' f f-J-, J ƒ " ƒf +
df K* e >_ dt ry?

which are valid along the characteristic lines given by dx


^ _= ±Cf and

dv do, 2 dii dv 2 T„, 2T (14)


v ^dt + C,G,— + p,R
— = 0
dt ' ' dt ' ' 'dt ' K* e

692 JOURNAL DE RECHERCHES HYDRAULIQUES, VOL. 32, 1994, NO. 5


which are valid along the characteristic lines given by 4y = ±C,, where

2v
Gf = c
P,( r -"')

G -
2v
"/ 2
P, ( c , - a2,)

The four compatibility equations, (Eqs 13 and 14), are integrated along the corresponding charac­
teristic lines. The time-line interpolation proposed by Goldberg and Wylie (1983) is used to solve
for p, v, ox, ü .
Downloaded by [North Carolina State University] at 16:22 30 November 2014

The shear stress between the fluid and the pipe wall, T„, is computed from the following equations:

K = *„, + X„„ (15)

Xm and Xwu are the steady and the unsteady shear stresses at the wall respectively. Expressions for
these stresses are

T,„ = pffU\U\/& (16)

T„„ = 2Pfvf(Yl + Y2+Y3)/R (17)

where

Y,(t+At) = Yl(t)e"^/"Umi[U(t + At) -1/(01 (18)

vf= kinematic viscosity; n, = 26.4, 200.0, 8000.0; m, = 1.0, 8.1, 40.0; for i = 1,2,3; and U= x>- ü.
A first-order approximation is used in Equation 18 where the velocity is set equal to the previous
time-step value.

Boundary Conditions
For the pipeline presented in Figure 1, the valve is located at the junction of two series pipes and
connected upstream to a tank which is maintained at a constant pressure head level, Hur. At the
downstream end the system is connected to another constant-head reservoir, Hdr. In the numerical
simulations, only the section from the upstream reservoir to the intermediate valve was modelled.
The section from the valve to the downstream tank includes the flowmeter devices, a long-radius
90° bend and a plexiglass pipe section for visualization. Instead of modelling this section the pres­
sure head history downstream of the valve, Hdxi, was recorded and used at the boundary.
At the upstream reservoir, the constant pressure is given by

p = ynur (i9)

JOURNAL OF HYDRAULIC RESEARCH, VOL. 32, 1994, NO. 5 693


where y = specific weight of the fluid. With the pressure known, the flow velocity in the two-equa­
tions model is obtained from the negative characteristic equation. For the intermediate valve, the
effective valve opening, denoted by x, and the pressure history downstream of the valve, Hdv, are
available. At any time during the simulation, the flow discharge, Q,n,, and the pressure at the
upstream side of the valve, ƒƒ„,., can be obtained by solving the positive characteristic equation and
the orifice equation simultaneously. The orifice equation can be expressed as:

Ö„„ = -j==jHuv-Hdu (20)

where, AH0 = head loss at the valve for the initial steady state, and Q0 = initial steady state dis­
Downloaded by [North Carolina State University] at 16:22 30 November 2014

charge.

MgmtÓtd Valve

C3)/®
4
Gait voivt
1 f \££

Fig. 1. Schematic of the Pipeline System.


For the FSI model, the fluid pressure at the upstream reservoir is computed from Equation 19. The
axial pipe velocity at the upstream end for a restrained pipe is given by

ü = 0 (21)

Equations 19 and 21 are solved with the two negative characteristic equations to determine Gv and x>
at the tank.

At the downstream end, the valve is fixed to the wall and the orifice equation is valid. Therefore
Equations 20 and 21 are solved together with the two positive characteristic equations to obtain p
and a, at the valve.

Description of the Experimental Apparatus


The experimental setup consists of two constant-head tanks open to the atmosphere and a motor­
ized butterfly valve located 32.0 m from the upstream tank and 12.74 m from the downstream tank.
The system is connected by a horizontal commercial steel pipe supported to the wall. The pipe
internal diameter and thickness are 0.101 m and 0.00635 m respectively. A pump supplies water to
the tank. The system is shown schematically in Figure 1. There is a short section of pipe, 0.30 m
long, extending vertically from the bottom of the upstream tank to a 90° elbow. After the elbow the
pipe is horizontal. The valve is flanged to the pipe by a short PVC connection. The total length of
the connection, including the valve itself is 0.80 m.

694 JOURNAL DE RECHERCHES HYDRAULIQUES, VOL. 32, 1994, NO. 5


The motorized device is a Grinnell butterfly valve coupled to a 24 DC motor by a chain-sprocket
system. Figure 2 shows the layout of the motor-valve coupling. The valve is activated with an
electric switch. Several modifications were tried to stop the valve smoothly at a specific angle.
However, even with highly deformable rubber cushions, the valve vibrated at the end of the run.
This was critical for the measurement of the valve opening angle during the valve motion. The
problem was solved by turning off the motor power using the electric switch and allowing the
valve to stop after the inertia effects were dissipated. This did not permit an exact prediction of
the final valve opening angle, but improved the stopping process and did not cause loss in the
accuracy of the measurements. Unfortunately, this prevented the exact repeatability of the
experiment.

Pointer
Downloaded by [North Carolina State University] at 16:22 30 November 2014

Fig. 2. Layout of the Motor-Valve Coupling System (a) Sectional view, (b) Plan view.

The discharge was measured for the initial and final steady state conditions. A thin square-edged
orifice (orifice to pipe diameter ratio = 0.69) and a 60° V-notch were used to measure the dis­
charge. The orifice satisfy the standards of the ASME as reported in Bean, (1971). Both instru­
ments gave practically the same discharge in the range between 0.0125 m3/s and 0.0015 m-Vs.
The maximum difference in the flow computations was 0.0001 mVs. Piezometric tubes were
installed along the pipe to obtain the pressure gradient during the steady state conditions. The
time history of the valve opening angle was obtained using a linear-potentiometer. The voltage
from the potentiometer was converted to degrees using a protractor capable of reading differ­
ences in angles up to 1.25°.

Prior to the transient experiments, a steady state calibration was carefully done to obtain the valve
loss coefficients at different opening angles. The head loss at the valve was measured from piezo­
metric tubes located at 4.8 and 3.14 diameters upstream and downstream of the valve. A total of 32
points were obtained for the valve loss coefficients and 56 points for the potenciometer calibration.
Careful attention must be given to this calibration in order to reproduce the valve characteristics
and its motion accurately. Watts et. al , (1980), reported accuracy problems in their experiments
near the valve fully closed position. The valve loss coefficients are presented in Figure 3 (0 vs \lk)

JOURNAL OF HYDRAULIC RESEARCH, VOL. 32, 1994, NO. 5 695


where an angle of zero degree corresponds to the fully open position. The head loss coefficient
through the valve is given by

*= 24(Hm-HdJ (22)
v'
where, v is the average velocity in the pipe and Hm and Hdv are the pressure heads upstream and
downstream of the valve respectively. As mentioned before, the accuracy of the calibration is par-
ticulary sensitive near the valve fully closed position. In this region, the loss coefficient, as defined
by Equation 22 tends to infinity.
Strain gage pressure transducers (Sensotec, Model TJE with an accuracy of 0.1%) were located at
10.88 m, 21.32 m (approx. 1/3 and 2/3 of the pipe length upstream of the the valve) and at both
Downloaded by [North Carolina State University] at 16:22 30 November 2014

sides of the valve. Their positions are shown in Figure 1. The transducers for the valve were
located at the same position as the piezometric tubes and are designed to record absolute pressures.
The signal was amplified using a Sensotec SA-BII Multichannel Amplifier and recorded on a PC
computer using a Computerscope (ISC-16) Data Acquisition System. The pressure transducers and
the potentiometer were triggered at the moment when the electric switch for the motor power was
activated. Computer software was developed for the conversion of voltages into water head and for
the generation of the valve closure curve as described in the next section. The use of a partial clo­
sure of an intermedite valve to generate a transient condition is a significant difference between this
and previous experiments reported in the literature.

lE+oi^— —I

IE+00

IE-01

7—

1E-02

1E-03

1E-04-I- — i — — i — — i — — i — — i — — i —
0 10 20 30 40 50 60 70 80
Valve Angle, degrees
Fig. 3. Valve loss coefficient vs. valve angle.

Description of the Valve Characteristics


The valve characteristics during a transient condition is commonly modelled by using a non-dimen-
tional parameter called effective valve opening, T, defined as (Chaudhry,1987 and Wylie and
Streeter, 1993):

696 JOURNAL DE RECHERCHES HYDRAULIQUES. VOL. 32, 1994, NO. 5


(C.A) (23)
(CA),
where Cd= discharge coefficient and Av = area of valve openings. The subscript o indicates refer­
ence conditions. A series of data interpolations were done to obtain the non-dimensional valve clo­
sure, x, versus time, t, relation. The voltage from the potentiometer gave a trace of the valve angle
position during the time of closure i.e. for any voltage measurement at time t, the valve opening
angle is known. The energy loss coefficients obtained from the steady state data at different valve
angles were assumed to be valid during the transient state. This is equivalent to assume that the ori­
fice equation for an incompressible flow applies under these conditions. Therefore, the loss coef­
ficient is known for any valve angle. Combining the valve angle versus time curve with the angle
Downloaded by [North Carolina State University] at 16:22 30 November 2014

versus loss coefficient curve it is possible to obtain the nondimensional valve opening, x, at any
time, t, from the following relation:
1 (24)
=i
where k„ was taken as the loss coefficient for the initial steady state condition and k, is the loss coef­
ficient at time t. It can be shown that Equation 23 is the same as Equation 24. Linear interpolation
was used to obtain the angle from the potentiometer data and parabolic interpolation was used to
obtain the loss coefficient from the interpolated angle. Figure 4 shows a typical X vs t curve com­
puted following this procedure. Smooth and continuous X vs t curves were obtained for partial clo­
sures from fully open position to maximum closure angle of about 72°. It must be stressed that, the
type of valve used has a thick rubber seal around the interior section that actually reduces the flow
to almost zero before the 90° angle is reached. It was observed that, after 83° the flow is com­
pletely stopped. At 72° the steady state valve discharge is 0.0017 m3/s i.e. 13.6% of the maximum
possible discharge.

1.0-1 -S^ 1

0.9 \

0.8- \

0.7 \

0.6 \

0.5- \

0.4- \

0.3 \.

0.2 N.

0.1 ^ ^ ^ ^

0.0-I 1 1 1 1 1
0 100 200 300 400 500 600
Time, ms
Fig. 4. Effective valve opening curve.

JOURNAL OF HYDRAULIC RESEARCH, VOL. 32, 1994, NO. 5 697


In general, the valve closure curve depends on the particular type of valve and the motor charac­
teristics. However, Figure 4 can be divided into three regions likely to occur in motor-regulated
valves. In region 1 the motor torque starts accelerating the valve from rest. The x - t curve
shows an increase in the slope until the valve reaches an almost constant speed. In Figure 4 this
region can be identified from x = 1 to X = 0.9 which occurs in a period of about 114 ms. Note
that the valve remained at rest for about 70 ms before the motor torque surpassed the valve
resistance.

Region 2 is characterized by a linear relationship between x and t. In Figure 4 this region goes from
X = 0.9 to x = 0.45 in about 100 ms. The value of X reduced more than 50% in this short period but,
interestingly enough, this is associated with a change in the closing angle of only 11° (from 0 = 8°
at x = 0.9 to x = 19° at x =0.45). Figure 8a shows that for regions 1 and 2 (a total of 220 ms) the
Downloaded by [North Carolina State University] at 16:22 30 November 2014

pressure at the upstream side of the valve remained practically constant. This is in agreement with
the small change in the valve position.
The last region, 3, is characterized by an exponential decrease in the x value. The slope of the curve
in this region has a major effect on the pressure rise history at the valve. The valve stops at about
520 ms where x reduces to a constant value of 0.029, (Figure 4). The first peak in pressure at the
upstream side of the valve occurs at this moment, (Figure 8a). Detailed information of the valve
position and the head loss coefficients in region 3 is recommended for accurate simulation of valve
closures in transient flow. Unfortunately, this data is not widely available.

Calculation of the Wave Speed


A procedure outlined by Martin (1983) for the calculation of the wave speed was adopted in this
research. The motorized valve was replaced by a butterfly valve with a mechanical arm. A very
rapid closure was generated by a fast movement of the arm. The closure time was less than half
of the theoretical period of the system. The pressure oscillations were recorded at position 3
(see Fig. 1) and a frequency analysis was done to obtain the power spectrum of these oscilla­
tions. The data sampling frequency was 1000 Hz. Figure 5(a) shows the pressure head history
for a sudden valve closure and Figure 5(b) shows its power spectrum. The frequency analysis
revealed a well-determined harmonic corresponding to a frequency of 8.79 Hz. The natural fre­
quency, F, for the pipe section from the upstream tank to the valve, assuming a single-phase
flow is

where af is the fluid wave speed and L is the pipe length between the valve and the upstream tank.
For L = 32.0 m the wave speed is 1125 m/s. The test was repeated for three different sampling fre­
quencies (200 Hz, 333.33 Hz and 1000 Hz) and the calculated wave speeds ranges from 984 m/s to
1125 m/s. A theoretical value of wave speed for pure water is 1300 m/s. The oxygen concentration
in the tanks was measured as 8.5 mg/1 indicating the presence of dissolved air in the water. This
explains the reduction in the wave speed.

698 JOURNAL DE RECHERCHES HYDRAULIQUES, VOL. 32, 1994, NO. 5


Downloaded by [North Carolina State University] at 16:22 30 November 2014

O 100 200 300 400 500 600 700 800 900 1000
Time, msec

Frequency, Hz
Fig. 5. Sudden valve closure data (a) Pressure history, (b) Power spectrum.

Results
The transient conditions generated by the motorized valve in the piping system, shown in Figure 1,
were simulated using the numerical models described in a previous section. The pressure history
downstream of the valve was measured by transducer number 4, (i.e Hdv) with a sampling rate of
2.0 ms, (Figure 6). The minimum pressure head recorded in this test is -6.7 m. The x - 1 curve and
the pressure history downstream of the valve (Figures 4 and 6) are used as an input to the models.
To avoid interpolation in any of the data input files, At was chosen as 2.0 ms and the selected wave
speed in the fluid for computation was adjusted to 900 m/sec. The measured pressure heads at sec­
tions 1, 2 and 3, (see Figure 1), are used to compare with the numerical results.
The wave speed in the pipe wall was used as 3 times the wave speed in the fluid to avoid interpola­
tion in the x -1 plane. The steady state flow was measured as 0.0125 mVs and the friction factor,

JOURNAL OF HYDRAULIC RESEARCH, VOL. 32, 1994, NO. 5 699


ƒ = .0225. The pressure head at the upstream reservoir, ƒƒ,„, was 2.5 m and the head loss in the
valve for initial discharge was measured as AH„ = 0.05 m.

10

5
I
\ i h A/\
Downloaded by [North Carolina State University] at 16:22 30 November 2014

-5
VJ \J V v^
10
0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.1
Time, sec
Fig. 6. Pressure history downstraeam of the valve.
The results for the two-equations model are presented in Figures 7a, 7b and 7c. Figure 7a presents
measured and simulated results at section 3, (at the valve). The comparison is excellent for the
entire simulation. Only a slight delay in the simulated results is observed in the second and the third
peaks. Figures 7b and 7c, shows the computed and measured pressure heads from transducers
located at 2L/3 and L/3 from the upstream tank respectively. The computed and measured first and
the second peaks compare satisfactorily. The general trends is also identified in the numerical
results, however; the presence of oscillations and reflections is noticed after 0.7 ms. The amplitude
of the computed oscillations is larger than the measured values.
The results for the FSI, with the frequency dependent friction model, are shown in Figures 8a, 8b,
and 8c. Figure 8a presents the measured and simulated results at section 3. Similar to the two-
equations model, the results at the valve show a very good agreement. Figures 8b and 8c show a
better numerical results than the first model and the higher frequency component of the pressure
wave is reduced giving a better overall simulation. There is some delay in the pressure history but
the damping rate is improved. A significant reduction in the first negative pressure peak is
observed in Figure 8b, when compared with Figure 7b.
For the simulations with the FSI model, the Young modulus of elasticity, e, was reduced to account
for the presence of the PVC pipe section located near the valve. There is some evidence that this
small part could act as a damping mechanism, (Ghilardi and Paoletti, 1986). Even though the pipe
was fixed to the wall, a small movement was noticed during the experiment. No measurements of
the pipe velocity or stress were taken, therefore it is impossible to quantify the effects of these
uncertainties.
When the frequency dependent friction effect is added to the FSI model, the differences can only
be noticed at the peak values. Figure 9 presents the results of the FSI without frequency dependent
friction at section 3. In Figure 9 a difference of about 0.5m is noticed between the numerical and
the measured results. This model has a stronger effect for fluids with higher viscosity.

700 JOURNAL DE RECHERCHES HYDRAUUQUES, VOL. 32, 1994, NO. 5


25:
Measured Values
Computed Values

15

•a
ra

I
Downloaded by [North Carolina State University] at 16:22 30 November 2014

-10 1
0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.1
Time, sec

JOURNAL OF HYDRAULIC RESEARCH, VOL. 32, 1994, NO. 5 701


25
- Measured Values
Computed Values
20!

15
Downloaded by [North Carolina State University] at 16:22 30 November 2014

-10
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 1.1
Time, sec

Fig. 7. Computed and measured results for sections 3, 2, and 1 using two equations model.

25
Measured Values

Computed Values

-10
0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 1.1
a. Time, sec

702 JOURNAL DE RECHERCHES HYDRAULIQUES, VOL. 32, 1994, NO. 5


25
- Measured Values

Computed Values
20

15

■a
ra
01
X
0)

I/>
V)
0)
Downloaded by [North Carolina State University] at 16:22 30 November 2014

-5

-10
0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.1
Time, sec

25
Measured Values

Computed Values

-5

-10
0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 1.1
Time, sec

Fig. 8. Computed and measured results for sections 3, 2 and 1 using FSI with frequecy dependent friction.

JOURNAL OF HYDRAULIC RESEARCH, VOL. 32, 1994. NO. 5 703


25
Measured Values

Computed Values

•a
ra
a>
I

01
Downloaded by [North Carolina State University] at 16:22 30 November 2014

-10
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 1.1
Time, sec

Fig. 9. Computed and measured results for section 3 using FSI and frequency-dependent friction.

Summary and Conclusion

Based on the experimental test and the numerical simulations, the following observation can be
made:

1. Either the two-equations or the FSI formulation can reproduce the pressure history at the valve
accurately, even for the smaller amplitude oscillations. This confirms that the measured pressure
head data and the computed x — t curve can be used to model the boundary condition imposed by
a partial-closure of an intermediate valve.
2. The two-equations formulation is adequate if the purpose of the computations is to determine the
first maximum positive pressure heads. After the second peak, the amplitude of the oscillations
was higher than the measured and low frequency oscillations appeared in the simulations.
3. The FSI formulation gave overall better numerical results. When the interaction between the
pipe and the fluid is modelled through Poisson coupling the damping of the oscillation is
improved considerably. Another improvement can be achieved when the frequency dependent
friction is added to the FSI model.

Acknowledgements

This work was supported in part by the National Science Foundation, through grant No. INT-9107009.

References

BEAN, H. S., 1971, Fluid Meters: Their Theory and Application, 6th ed., ASME publication, New York, NY.
BUDNY, D. D., WlGGERT, D. C, and HATFIELD, F. J., 1990, "Energy Dissipation in the Axially Coupled Model
for Transient Flow", BHRA Pressure Surges-Proceedings of the 6th International Conference, pp. 15-26.

704 JOURNAL DE RECHERCHES HYDRAULIQUES, VOL. 32, 1994. NO. 5


CHAUDHRY, M. H., 1987, Applied Hydraulic Transients, 2nd ed., Van Nostrand Reinhold, New York, NY.
FORSYTHE, G. E., and WASOW, W. R. , I960, Finite-Difference Methods For Differential Equations, John
Wiely & Sons, Inc.
GHILARDI, P. and PAOLETTl, A., 1986, "Additional Viscoelastic Pipes as Pressure Surges Suppressors",
BHRA Pressure Surges-Proceeding of the 5th International Conference, pp. 113-121.
GOLDBERG, D. E. and WYLIE, E. B„ 1983, "Characteristics Method Using Time-Line Interpolation", ASCE
Journal of Hydraulics Division, Vol. 109, No. 5.
MARTIN, C. S., 1983, "Experimental Investigation of Column Separation with Rapid Closure of Downstream
Valve", BHRA Proceedings of the 4th. International Conference on Pressure Surges, England.
Suo L. and WYLIE E. B., 1989, "Impulse Response Method for Frequency Dependent Pipeline Transients",
ASME Journal of Fluid Engineering, Vol. 111, pp. 478-483.
TRIKHA, A. K., 1975, "An Efficient Method for Simulating Frequency-Dependent Friction in Liquid Flow",
ASME Journal of Fluids Engineering, pp. 97-105.
VARDY, A., 1992, "Approximating Unsteady Friction at High Reynolds Numbers", Proceeding of the Interna-
tional Conference on Unsteady Flow and Fluid Transients, Durham, United Kingdom, A.A.Balkema,
Rotterdam.
VARDY, A.E. and HWANG K., 1991, "A Characteristics Model of Transient Friction", Journal of Hydraulics
Downloaded by [North Carolina State University] at 16:22 30 November 2014

Research, Vol. 29, No. 5, pp. 669-684.


WATT, C. S., BOLDY, A. P., HOBBS, J. M., 1980, "Combination of Finite Difference and Finite Element Tech­
niques in Hydraulic Transient Problems", BHRA, Proceedings of the 3rd. International Conference on
Pressure Surges, England.
WlGGERT, D. C , OTWELL, R. S., and HATFIELD, F. J., 1985, "The Effect of Elbow Restraints on Pressure
Transients", ASME Journal of Fluids Engineering, Vol. 107.
WIGGERT, D. C, HATFIELD, F. J. and STUCKENBRUCK, S., 1987, "Analysis of Liquid and Structural Transients
by the Method of Characteristics", ASME Journal of Fluid Engineering, Vol. 109, pp. 161-165.
WYLIE, E. B., and STREETER, V. L., 1993, Fluid Transients in Systems, Prentice-Hall, Englewood Cliffs, NJ.
ZlELKE, W., 1968, "Frequency Dependent Friction in Transient Pipe Flow", ASME Journal of Basic
Engineering, Vol. 90, No. 1, pp. 109-115.

Notations:

A Cross-sectional area, m2
Av Opening area of valve, m2
a Speed of sound, m2
cd Orifice discharge coefficient
c Coupled wave speed, m/sec
D Pipe diameter, m
E Young's modulus of elasticity of pipe wall material, Pa
e Pipe wall' thickness, m
F Natural frequency, Hz
f Darcy-Weisbach friction factor.
G Parameter in the FSI formulation,
H Pressure head, m
AH Head loss at the valve, m
K Bulk modulus of fluid, Pa
k loss coefficient,
L Pipe length, m
P Pressure, Pa
R Pipe inside radius, m
t Time, sec
U Relative fluid velocity, (v - u), m/sec
u Pipe axial velocity, m/sec
V Fluid velocity, m/sec
X Axial coordinate

JOURNAL OF HYDRAULIC RESEARCH, VOL. 32, 1994, NO. 5 705


y, Parameter in the F D F
$ Specific weight, = pg
v Poisson ratio
\if Kinematic viscosity, m 2 sec
p Density, kg/m 3
a Pipe stress, Pa
x Effective valve opening,
xw Wall shear stress

Subscripts
dr Downstream reservoir
dv Downstream of the valve
Downloaded by [North Carolina State University] at 16:22 30 November 2014

ƒ liquid
5 Steady state conditioner
t Pipe
ur Upstream reservoir
uv Uptream of the valve
u Unsteady-state
x Axial direction
o Initial condition

706 JOURNAL DE RECHERCHES HYDRAULIQUES, VOL. 32, 1994, NO 5

You might also like