Sei sulla pagina 1di 16

Water Resour Manage

DOI 10.1007/s11269-016-1310-1

Study of a Compressed Air Vessel for Controlling


the Pressure Surge in Water Networks: CFD
and Experimental Analysis

Mohsen Besharat 1,2 & Reza Tarinejad 1 &


Mohammad Taghi Aalami 1 & Helena M. Ramos 2

Received: 8 June 2015 / Accepted: 28 March 2016


# Springer Science+Business Media Dordrecht 2016

Abstract Air chambers show good ability in controlling the pressure surge from a water
hammer (WH) phenomenon. To simulate an air chamber and study the behavior of air
inside it, a compressed air vessel (CAV) is considered in a pressurized system. The
current work consists of experimental tests and one-dimensional (1D) and two-
dimensional (2D) computational fluid dynamics (CFD) simulations for an air pocket
within a CAV in the case of rapid pressurization and the occurrence of WH in a
pressurized system. The pressure variations create vorticity and turbulence with oscillat-
ing behaviors, but the available 1D models are unable to simulate those phenomena
adequately. Therefore, by using the measured data, proper CFD analysis is conducted
considering the effect of the wall, y+, mesh size, turbulence, and the wall treatment
method to better understand the behavior of the system. Results of the CFD simulation
show that realizable k-ε turbulence model, when coupled with the enhanced wall
treatment (EWT) method, works adequately for modeling the pressure oscillation. The
volume of fluid (VOF) model and the piecewise linear interface calculation (PLIC)
method have presented good ability in the prediction of the air-water interface.

Keywords Water hammer . CFD . Pressure surge . Air vessel . Pressure oscillation

Abbreviations
a Pressure wave speed
g Gravitational acceleration

* Reza Tarinejad
r_tarinejad@tabrizu.ac.ir

1
Faculty of Civil Engineering, University of Tabriz, 29 Bahman Boulevard, East Azerbaijan, Tabriz,
Iran
2
Department of Civil Engineering and Architecture, Instituto Superior Técnico (IST), University of
Lisbon, Avenida Rovisco Pais, 1, 1049-001 Lisbon, Portugal
M. Besharat et al.

ρ Density
γ Water specific weight
υ Kinematic viscosity
μ Dynamic viscosity
μt Turbulent dynamic viscosity
V Mean Velocity
H Head
t Temporal coordinate
x y, Spatial coordinates
τw Wall shear stress
u Velocity component
P∗ao Absolute pressure within air pocket
y+ Dimensionless distance from the wall parameter
y∘ Distance from the wall
u* Friction velocity
u+ Dimensionless velocity parameter
k Turbulence kinetic energy
ε Dissipation rate
Sij Strain rate
S Mean strain rate
Ret,y Turbulent Reynolds number
λC Specific variable of cell C
aC Center coefficient
aNeighbor Coefficient of neighborhood cells
b Constant part of source term and boundary conditions
RλUSC Unscaled residual
RλSC Scaled residuals
ωi Fluctuating vorticity
ωk Angular velocity
Ωi Mean vorticity
Ωij Rotation rate
Ωi j Mean rotation rate
κ Kármán constant
Γ Blending function

Subscripts
lam Laminar
tur Turbulent
∘ Initial value

Acronyms
HT Hydro-pneumatic Tanks
PT Pressure Transducers
RVP Rapid Void Pressurization
CAV Compressed Air Vessel
WH Water Hammer
Study of a Compressed Air Vessel for Controlling the Pressure

1 Introduction

1.1 Single-Phase Transient Flow

Unsteady flow refers to flow in which the pressure and velocity vary with time. In many cases,
transient flow is associated with pressure surges and high-velocity waves. While engineers
design water networks to operate predominantly at steady state conditions, the transient state
should be considered since it can result in significant pressure oscillations that may exceed the
design limits of pipes and fittings. The efforts of various researchers, e.g., Jaeger (1933), Wood
(1937), Rich (1944), Rich (1951), Parmakian (1955), Streeter and Lai (1963), and Streeter and
Wylie (1967), led to the classical mass and momentum equations for one-dimensional (1D)
hydraulic transient flow, i.e.,:

a2 ∂V ∂H
þ ¼0 ð1Þ
g ∂x ∂t

∂V ∂H 4
þg þ τw ¼ 0 ð2Þ
∂t ∂x ρD
where V is the mean velocity, H is the head, t is the temporal coordinate, x is the spatial
coordinate along the pipe, and τw is the shear stress at the wall. The most popular method for
solving hydraulic transient problems using a finite difference procedure in the 1D approach is
the well-known method of characteristics (MOC) (Wylie 1993). The MOC is more accurate
than other methods, particularly for a constant wave speed. It is convenient for capturing the
location of the steep wave fronts, the illustration of the wave propagation, easy programming
and computation (Chaudhry 2014). The Bentley Hammer model is a well-known software
package that is used in this research to analyze 1D hydraulic systems by the MOC technique.

1.2 Two-Phase Transient Flow

The presence of air in hydraulic systems changes the behavior of the flow, that MOC is unable
to predict characteristics of the air precisely. Most existing models ignore the air, and other
models incorporate air-water interactions. Lee (1994) provided an explanation of the under-
estimation of the peak pressure by standard WH theory and they showed the pressure could be
significantly higher than the pressure predicted by the standard theory with a fixed-speed
wave. Falk (2000) reported a modified MOC that works well with pressure waves, but it
cannot predict void waves and it underestimates the peak pressure. Generally, to simplify the
problem, it is assumed that the air pocket deforms uniformly with no spatial variation in the gas
region. Therefore, the propagation of the pressure wave can be neglected in the gas phase
(Chaiko and Brinckman 2002). These issues indicate the need for a multi-phase flow model.

1.3 Transient Shear Stress on the Wall

Unsteady flows in pipe networks usually are analyzed by means of 1D models, in which the
energy dissipation is based on the friction factor, which in turn, depends on the local and
instantaneous Reynolds number (quasi-steady model). Such models underestimate the friction
M. Besharat et al.

forces and overestimate the persistence of oscillations following the first peak. Indeed, unsteady
friction is of great importance in pipe networks, and it should be considered precisely because
oscillation of the maximum pressure can occur after the first cycle due to reflexion effects.
Ramos et al. (2004) presented dimensionless equations of pressurized transient flows to develop
a generic formulation applicable for any system. The authors presented two corrective coeffi-
cients for cases in which there were no velocity profile data. This method improved the
agreement between numerical and experimental data in terms of damping and phase shift.
Kwon (2007) compared the results of 1D and 2D models by investigating the friction effect in
MOC and the axisymmetrical 2D model. The author concluded that the friction coefficient used
for MOC was much higher than the Darcy-Weisbach friction coefficient and that the κ parameter
used for estimating the mixing length in the axisymmetrical model should be smaller.
The motivation for conducting this research consists of the need for more progress in
the analysis of pressurized transient flow containing air. Although the actions that cause
strong pressure variations are well known, to date, the effects of pressure damping and
propagation are not well understood (Ramos et al. 2004). To address this factor, the WH
phenomenon was analyzed experimentally, and 1D and 2D CFD simulations of an
integrated pressurized system were prepared. The considered CAV could be used as an
air chamber to control pressure surges. The characteristics of such an air pocket have the
potential for providing significant benefits, and this is the reason current work focused
on studying the behavior of air pockets for pressurization induced by the WH.

2 Numerical Method and Boundary Conditions

The lack of understanding of the changes in turbulence during transient flow is a significant
obstacle to identify problems and to generate more accurate models (Ghidaoui et al. 2005). 2D
and 3D CFD models seem to be able to calculate pressure fluctuations accurately. However,
due to complexities in the governing equations, the use of CFD models has been always
challenging and time-consuming for two-phase transient states. Among 2D and 3D CFD
simulations, the 2D CFD simulations are more preferred due to less calculation time and effort.
In a comparison of 2D and 3D CFD simulations, Zhou et al. (2011) concluded that the
simulations of 2D and 3D were in good agreement, but the 2D is simpler and quicker.
Using correct values of shear stress in the CFD calculations has a significant effect on
the final results. CFD techniques are capable of calculating satisfactory values of shear
stresses if a fine mesh is generated or appropriate inflation layers are provided near the
walls. The widely-referenced parameter y+ = u* y∘/υ supplies a tool that helps to under-
stand the near-wall flow condition by dividing the boundary layer into several sub-layers.
An increased level of shear stress in the boundary layer occurs because of the diffusion
effect of the turbulence. The y+ relation can be used to calculate the height of the first
near-wall cell (y∘) based on a specific y+ value. In this work, an appropriate value of y+
is selected based on the turbulence model and wall-treatment method or function. This
calculation produces a good estimate of the extent to which the near-wall zones should
be inflated to obtain the best calculation of wall effect.
Using an appropriate turbulent model will increase the efficiency and accuracy of the simula-
tion. There are many turbulent models, ranging from one-equation models to robust, two-equation
models. Two-equation models are used extensively in fluid simulations (Mathur and He 2013).
Study of a Compressed Air Vessel for Controlling the Pressure

The standard two-equation k-ε model is a semi-empirical turbulence model based on the eddy
viscosity (turbulent-viscosity) concept (μt = ρCμk2/ε), and it has become known as a powerful,
economical, and reasonably accurate model for a wide range of turbulent flows. The Cμ in μt
relation is constant in the standard k-ε model. This model uses two transport equations for
turbulence kinetic energy (k) and its dissipation rate (ε). The transport equation for k is derived
from exact formulations while the dissipation rate (ε) is derived from an empirical formulation
(Launder and Spalding 1972; Cebeci 2004). The standard k-ε model works quite well for boundary
layer flows, but it has some deficiencies in simulating massive separation, calculating the length
scale for turbulent flows and flows in which the pressure has large fluctuations. The realizable
k-ε model works better in these cases by taking advantage of improvements in the
calculations of μt and ε. It is known that μt formulation becomes non-realizable in the
case of large mean strain rate. To avoid that, the Cμ constant in μt formulation must be
dependent on the mean strain rate (S) rather than being constant. In addition, an exact
formulation for the ε transport equation was used in the realizable k-ε model by
developing a model for the mean-square vorticity fluctuation, i.e., ωi ωi (Shih et al.
1995). In a two-phase flow simulation work, Dostal et al. (2008) used the realizable
k-ε model to simulate the two-phase flow of Pb-Bi and steam by means of Fluent
analysis package. The realizable k-ε model showed good ability in the simulation of
the bubbly situation and formation of initial bullet shape bubble in slug flow with no
significant increase in computational time. Martins et al. (2014) used the realizable k-ε
model in the Fluent environment to analyze velocity distribution in pressurized pipe flow
and achieved good results.
The realizable k-ε model is considered for turbulence modeling in current research. The
realizable k-ε model uses an improved transport equation for ε written as Eq. (3).
  
∂ε ∂ε ε2 ∂ μ ∂ε
ρ þ ρu j ¼ ρC 1 Sε−ρC 2 pffiffiffiffiffi þ μþ t ð3Þ
∂t ∂x j k þ υε ∂x j σε ∂x j

where k ≈ u2 and ε ¼ υðωi ωi Þ. The fluctuating vorticity (ωi) and S can be written as
ωi = εijkup and
k,jffiffiffiffiffiffiffiffiffiffiffiffiffi
S ¼ 2S i j S i j , respectively, as well as Sij = 0.5(ūi,j + ūj,i). In these formulations, in general,
the ui,j parameter is defined as ∂ui/∂xj. The values of the constants are C2 = 1.90, σε = 1.20.
Also, C1 is calculated by Eq. (4) (Shih et al. 1995).
 
Sk
C 1 ¼ max 0:43 ; ð4Þ
5ε þ S k

The Cμ parameter is not constant in the realizable k-ε model and it was calculated using Eq.
(5).
1
Cμ ¼ ð5Þ
kU *
Ao þ As
ε
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi pffiffiffi
where U * ¼ Si jSi j þ Ω ~ i jΩ~ i j ; Ao ¼ 4:04; As ¼ 6cosϕ, and ϕ is defined as:
" #
1 pffiffiffi S i j S jk S ki
ϕ ¼ arccos 6  3=2 ð6Þ
3 Si jSi j
M. Besharat et al.

The formulation for the mean vorticity (Ωi) was introduced by Shih et al. (1995) as
Ωi = εijkūk,j, and, from that, Ωij = 0.5(ūi,j − ūj,i) was defined. Shih et al. (1994) introduced Ω~ ij
* *
¼ Ω i j −2εi jk ωk and Ω i j ¼ Ωi j −εi jk ωk , where Ωi j is the mean rate of the rotation tensor
viewed in a moving reference frame with the angular velocity of ωk.
A major limitation of two-equation turbulent models is that their direct application in no-
slip wall behavior gives unsatisfactory results near the wall due to their inability to calculate
the constant, B, in the law of the wall equation, i.e., u = u*[ln(u*y/υ)/κ + B] where κ and B are
constants. This is a consequence of the presence of a very thin layer in which the μt changes
rapidly with distance from the wall (Mathur and He 2013). To overcome this defect, wall
functions are frequently proposed and used while this approach does not generate satisfactory
results for separated flows, moving parts, and high-pressure gradients (Wilcox 2006). The
enhanced wall treatment (EWT) method is addressed in these cases which is a method for
using in the ε that combines a two-layer model with enhanced wall functions (Zhang et al.
2013). This method is able to resolve the near-wall viscous sub-layer that usually means
having the first cell inside the y+ ≈ 1 region. However, the procedure requires a heavy
computational cost if the requirement of complete fine meshing near the wall is preserved.
The two-layer approach is used here to determine ε and μt near the wall by dividing the
domain into two regions, i.e., a viscous sub-layer and a turbulent sub-layer, which are
pffiffiffi
separated by a wall distance-based named turbulent Reynolds number, i.e., Ret;y ¼ ρy k =μ
for Ret,y < 200 as the viscous sub-layer and Ret,y > 200 as turbulent sub-layer. Zhang et al.
(2013) compared the adjusted wall function, standard wall function and EWT for modeling the
indoor convective heat transfer and flow in a mixing ventilation mode and an under-floor
displacement ventilation mode using the Fluent software where the EWT showed appropriate
performance.
The EWT method is used in this work to calculate the specifications of the flow in the near-
wall region. This method combines the linear and logarithmic laws of the wall by utilizing a
function from Kader (1993), considering laminar flow for the linear law of the wall and
turbulent flow for the logarithmic law of the wall:

u þ ¼ eΓ u þ þ
1
lam þ e utur ð7Þ
Γ

where Γ is the blending function defined by Γ = 0.01(y+)4/(1 + 5y+).


To date, many methods have been proposed to determine the deformation of the interface in
the simulation of a multi-phase flow. Three methods are mainly deployed, i.e., the volume of
fluid (VOF), level-set, and front-tracking (Wörner 2003). Among the three, the VOF model is
used extensively because it offers a straightforward and economical procedure for tracking
interfaces with a minimum storage requirement. Zhou et al. (2011) showed that the VOF
model is effective for the simulation of air pocket movement and the resultant pressure surge in
the simulation of a rapid filling pipe containing entrapped air using the Fluent software. The
VOF model determines the fractional volume line inside a cell which is known as reconstruc-
tion. Reconstruction is normally done either by the simple line interface calculation (SLIC)
method or the piecewise linear interface calculation (PLIC) method (Wörner 2003). Due to its
higher accuracy, the PLIC method is used in the current study to specify the position of the
interface.
The applied boundary conditions for HT1 and HT2 are shown in Figs 1 and 2. All the other
parts of the model are defined as no-slip wall boundary.
Study of a Compressed Air Vessel for Controlling the Pressure

3 Research Outline

An apparatus (Figs 1 and 2) that included two hydro-pneumatic tanks (HT) (each 1 m3 in
volume) and connected by a series of pipes (about 8 m in length) having nominal
diameters of 63 mm (DN63) and a nominal pressure of 16 bars (PN16) was used to
investigate the effect of pressurization of an air pocket. There were four ball valves with
electronic pneumatic actuators identified as V1, V2, V3, and V4. Three pressure trans-
ducers were installed at different locations, identified as PT1, PT2, and PT3. A trigger was
used to electrically activate all measuring equipment and to synchronize the
measurements. Martins (2013) first established the base system with one HT and a
vertical pipe, as shown in Fig. 1a, to study the rapid pressurization of an air pocket in
a vertical pipe (system 1). Later, the system was improved as shown in Fig. 1b (system
2). For the first set of tests, rapid void pressurization (RVP) under high pressure induced
by two HTs was fulfilled to study the rapid pressurization of a small air pocket at the
highest point of the pipe profile in the system 2 and CFD simulations were conducted to
model the operation of the system. First, the CFD model was developed to simulate
pressure variations in system 1. After the model was calibrated, its accuracy was
validated in system 2 (see section 4). This part was considered as a reference to obtain
the information required to conduct the CFD analysis in more challenging models that
contained moving cell zones resulting from the valve closure maneuvers.
For RVP test in system 2, a 0.25 cm in length air pocket was considered with an initial
pressure of 1 bar. During the RVP tests, the pressure in the air pocket reached a peak value of
7.25 bar. Then, it began to fluctuate and merge until a stable value of pressure was reached that
corresponded to the initial pressure of HT1, i.e., 3 bars.
System 3 (Fig. 2) provided an appropriate tool to investigate the behavior of air pocket in a
CAV with a diameter of 140 mm and height of 325 mm under WH test. Tests were conducted
with a WH event induced by the closure of valve V3. Then, the increases in the pressure in the
air pocket were measured. It takes 0.20 s time for the opening of the valve during the WH tests.

Fig. 1 a Installation and components of systems, (a) System 1 b Installation and components of systems, (b)
System 2
M. Besharat et al.

Fig. 2 a Installation of CAV in the WH test and components (system 3), (a) Schematic b Installation of CAV in
the WH test and components (system 3), (b) Real photo

4 Mesh Analysis and Validation of Numerical Results

Authors attempted to model every component of the physical models (orientation,


elbows, valves, CAV, HTs, and pipe branches). The geometry was created in 2D using
CAD tools, at which unstructured and structured meshes was generated, accurately
depicting all parts of the system. The 2D geometry was considered due to the size of
the apparatus and the huge calculation load associated with 3D geometry. Several
inflation layers were considered near the wall to calculate the exact values of shear
stress. The y+ relation was used to calculate the height of the first layer (y∘). Since EWT
method is supposed to be used in this model, for the first try, y+ ≈ 1 was tackled and,
based on that, the height of the first layer was calculated as approximately 1 × 10−4 m.
CFD simulations were done in the ANSYS Fluent environment. The scaled residual
monitoring (SRM) method was used to monitor convergence. The unscaled residuals
were scaled as described below. Considering the conservation equation of a specific
variable λ at cell C as:

 
∑ aNeighbor λNeighbor þ b−aC λC ¼ 0 ð8Þ

where aC is the center coefficient, aNeighbor is the coefficient of neighborhood cells, and b
is the constant part of the source term and boundary conditions, then the unscaled
residual (RλUSC) can be defined as:

all cells h i
X X 
RλU SC ¼ aNeighbor λNeighbor þ b−aC λC ð9Þ
Study of a Compressed Air Vessel for Controlling the Pressure

The scaled residuals for all variables (RλSC) were calculated by:

all cells h i
X X 
aNeighbor λNeighbor þ b−aC λC
RλSC ¼ all cells
ð10Þ
X
ðaC λC Þ

The convergence criterion for the residual values was selected as 1 × 10−8 for the energy
equation and as 1 × 10−5 for all of the other equations.
System 2 was calculated with different meshes to investigate the effect of the size of the
mesh and the number of cells on the final results using a mesh independent analysis (MIA).
During the generation of the mesh, the skewness and aspect ratio of the cells of the mesh
always were controlled to guarantee the quality of the cells (Table 1).
The VOF model and the PLIC geometry reconstruction scheme were used for multi-
phase modeling of the flow. Moreover, numerical solution techniques of pressure implicit
with the splitting of operators (PISO) for coupling the solutions of the pressure–velocity
equations were considered. In respect of the air pocket behavior, ideal gas law for
compressible flows was used to determine the behavior of the air pocket because the
Mach number was greater than unity.
The MIA began by generating unstructured meshes with triangular cells for the main body
and quadrilateral cells for the inflation layers (Fig. 3). The analysis was continued by
generating a structured mesh (Mesh 4), as shown in Fig. 3. An ordinary PC (Core 2, CPU
2.14 GHz, RAM 3 GB) was used for the calculations. The CFD model was calibrated by
modeling the system 1 which had been simulated in a previously published work of Martins
et al. 2012 as shown in Fig. 4. The root mean square errors (RMSE) for simulations are
presented in Fig. 4. Then, the CFD model was used for simulations in system 2. Figure 5
shows an acceptable agreement for all meshes with the experimental measurements of the
results of pressure oscillations. The best match was achieved for Meshes 1, 2, and 4. But, the
long time required to generate the meshes was a significant challenge to the efficiency of the

Table 1 Meshing information for different CFD 2D simulations system 2

Mesh 1 Mesh 2 Mesh 3 Mesh 4

Cell Type Triangular Triangular Triangular Quadrilateral


Number of Cells 260769 255399 90402 165537
Number of Faces 415499 407444 143581 333816
Number of Nodes 154731 152046 53180 168280
Maximum Face Area (m2) 0.0104 0.0104 0.0169 0.015
Maximum Skewness 0.77 0.77 0.54 0.64
Maximum Aspect ratio 42.49 14.09 7.38 10.84
Inflation Layer No. 10 10 5 Bias Inflation
y+ 5.5 7.6 10 - 20 15
First Layer Height (m) 0.00015 0.0005 0.0015 0.0015
Calculation Time (hour) 91.92 91.75 23.36 31.25
Time Step (s) 0.0001 0.0001 0.0001 0.0001
M. Besharat et al.

Fig. 3 a Different generated meshes: (a,b) quadrilateral mesh (mesh 4) b Different generated meshes: (a,b)
quadrilateral mesh (mesh 4) c Different generated meshes: (c,d) triangular mesh Fig. 3(d) Different generated
meshes: (c,d) triangular mesh

Fig. 4 a Validation of numerical results, (a) air pocket length of 0.5 m b Validation of numerical results, (b) air
pocket length of 0.1 m
Study of a Compressed Air Vessel for Controlling the Pressure

Fig. 5 Experimental and CFD simulation of pressure variation for different meshes at RVP

process. Several configurations of the meshes were tried to reduce the calculation time while
maintaining accuracy to the maximum extent possible, and Mesh 3 provided consistent
pressure results with much lower calculation time. The small shift between simulation and
experimental results may come from some inconsistencies between the numerical model and
the physical model, e.g., the large number of fittings and connections in the physical model.

5 Results and Discussion

Figure 6 presents the velocity streamline (in the air pocket and at the entrance of HT2) and the
volume fraction of both phases (air and water) from the simulation results of Mesh 1. The
results include the starting time (0.015 s), time that the highest velocity was attained (0.06 s),
time that the maximum pressure was attained (0.08 s), and the time for the air-water mixture
after the first peak (0.12 s). The air-water interface is deformed in several times, showing that
how the air pocket acts during a transient action. The streamlines indicate the changes in
velocity in two parts of the model and the flow characteristics associated with the direction of
displacement. The simulation detected a separation in the entrance of HT2 (e.g., time = 0.08 s),
where two elbows changed the pipe’s orientation.
Figure 7 shows good agreement for the pressure variation inside the air pocket during a WH
test in system 3. During all the tests, the initial pressures of the air pocket were kept in the
atmospheric range, i.e., 1 bar. CFD analyzes of the WH tests were conducted using a model that
was configured with a separate zone for simulating actuation by a valve. The valve zone was
completely separated from the main body by two interfaces, allowing the rotation of the body of
the valve. The angular velocity (ωk) of valve rotation was defined as 2.5π, similar to the
experimental situation in the definition of the valve’s zone rotation. Based on the information
obtained from the CFD simulation of system 2, a mesh similar to Mesh 3 was generated to earn
acceptable pressure results in the shortest calculation time. The information pertaining to the
generated mesh is presented in Table 2. The RMSEs for this simulation are presented in the Fig. 7.
M. Besharat et al.

Fig. 6 a Velocity Streamline (m/s) and air-water interface position in CFD 2D simulation of RVP test, (a,b) time
0.015 s b Velocity Streamline (m/s) and air-water interface position in CFD 2D simulation of RVP test, (a,b) time
0.015 s c Velocity Streamline (m/s) and air-water interface position in CFD 2D simulation of RVP test, (c,d) time
0.06 s d Velocity Streamline (m/s) and air-water interface position in CFD 2D simulation of RVP test, (c,d) time
0.06 s e Velocity Streamline (m/s) and air-water interface position in CFD 2D simulation of RVP test, (e,f) time
0.08 s (maximum pressure point) f Velocity Streamline (m/s) and air-water interface position in CFD 2D
simulation of RVP test, (e,f) time 0.08 s (maximum pressure point) g Velocity Streamline (m/s) and air-water
interface position in CFD 2D simulation of RVP test, (g,h) time 0.12 s h Velocity Streamline (m/s) and air-water
interface position in CFD 2D simulation of RVP test, (g,h) time 0.12 s

Looking at the nexus between the 1D and 2D models, a 1D analysis was conducted using
Bentley Hammer software (Fig. 7). Similar to the physical model, minor losses were defined in
every valve and change in geometry. In addition, friction losses were introduced by defining
the pipe material and the speed of the wave.
Study of a Compressed Air Vessel for Controlling the Pressure

Fig. 7 Pressure results of WH test, 1D simulation, 2D simulation, and experimental data within CAV

Figure 8 shows the velocity contours and the air-water interface during the WH tests.
The results demonstrate how the valve maneuver affected the flow propagation for
various positions such as CAV entrance, geometry change and valve. The flow was
moving toward the CAV when the test began until the pressure in the air pocket
increased and prevented any additional flow from entering the CAV. This was accom-
panied by a small elevation of the air-water interface. Then, the flow followed the path of
the pipe in the direction of HT2 (5.75 s). At the test time, after about 5.85 s, the valve
(V3) began to actuate closure. The contours show that the velocity increased inside the
body of the valve during the process of closing and the air pocket was compressed by the
movement of the air-water interface towards the top of the CAV. The valve was closed
completely at 6.05 s, and then, the opening actuation began.

Table 2 Meshing information of CFD 2D simulation of system 3

Mesh Information

Cell Type Triangular


Number of Cells 87134
Number of Faces 151484
Number of Nodes 64242
Maximum Face Area (m2) 0.03
Maximum Skewness 0.89
Maximum Aspect ratio 40
Inflation Layer No. 10
y+ 8
First Layer Height (m) 0.0003
Calculation Time (hour) 31.87
Time Step (s) started with 1 × 10−5 and continued with 1 × 10−4
M. Besharat et al.

Fig. 8 a Velocity contours (m/s) and air-water interface position at the vicinity of V3 b Velocity contours (m/s)
and air-water interface position at the vicinity of V3 and CAV of WH test (VFR = 15.84 %), (b) 5.85 s c Velocity
contours (m/s) and air-water interface position at the vicinity of V3 and CAV of WH test (VFR = 15.84 %), (c)
5.89 s d Velocity contours (m/s) and air-water interface position at the vicinity of V3 and CAV of WH test
(VFR = 15.84 %), (d) 5.93 s e Velocity contours (m/s) and air-water interface position at the vicinity of V3 and
CAV of WH test (VFR = 15.84 %), (e) 5.97 s f Velocity contours (m/s) and air-water interface position at the
vicinity of V3 and CAV of WH test (VFR = 15.84 %), (f) 6.01 s g Velocity contours (m/s) and air-water interface
position at the vicinity of V3 and CAVof WH test (VFR = 15.84 %), (g) 6.05 s h Velocity contours (m/s) and air-
water interface position at the vicinity of V3 and CAV of WH test (VFR = 15.84 %), (h) 6.15 s
Study of a Compressed Air Vessel for Controlling the Pressure

6 Conclusion

The research was conducted on transient two-phase flow in a pipe system with a CAV. A MIA
was conducted to find an efficient mesh. Table 1 clearly shows that a longer computation time
corresponded to a higher number of cells and a smaller height of the first layer. Mesh 3 was
defined to give quite acceptable results for the pressure variations while having much shorter
time for generating the mesh and performing the calculations. Mesh 4 was created using
manipulated quadrilateral cells, and despite the length of time required to generate the mesh, it
had advantages over the other meshes, including short calculation time and high accuracy. The
results of the CFD simulations prove the ability of the realizable k-ε turbulent model, when
used in conjunction with EWT method, to predict the results of a wide range of pressures.
Nonetheless, the 2D model was unable to generate pressure results mostly after the second and
third peaks of pressure waves. This was the consequence of accumulating errors in the
calculation of the flow specifications after several transient movements in different directions.
Some considerable problems that were addressed in 2D analyzes were long simulation time,
decidedly unstable calculation behavior, and the convergence problem. Because of using the
VOF model and having two phases, courant number was considered less than unity. Based on
that, a courant number equal to 0.2 was selected for simulations. The time step was calculated
using this courant number and minimum size of mesh element. The 2D CFD simulation of the
RVP was conducted with a time step of 1 × 10−4 s while the 2D simulation of WH test began
with a time step equivalent to 1 × 10−5 s for some periods and continued with 1 × 10−4 s until
the end. The transition from the time step 1 × 10−5 s to 1 × 10−4 s was fulfilled using a variable
time-stepping method. This method uses a function that observes the convergence criteria of
the simulation accurately in every time step. If the simulation is converging, i.e., the Courant
number is decreasing in each time step, the function increases the initial time step by a defined
increment value and vice-versa.
The results show that the 1D models cannot be used confidently to predict the pressure
changes of a transient flow with an air pocket. In addition, 1D analysis tools were unable to
predict the position of the air-water interface accurately, which will have a considerable role in
the behavior of an air pocket when it is being pressurized.

Acknowledgment The authors acknowledge the Civil Engineering, Research, and Innovation for Sustainability
(CEris) Center, Department of Civil Engineering, Architecture and Georesources, Instituto Superior Técnico
(IST), University of Lisbon (ULisbon), Portugal for providing the experimental facilities for conducting the tests
and financial support for the first author under grant number of HR Serviços no 2714.

Compliance with ethical standards

Conflict of Interest The authors declare that there are no conflicts of interest.

References

Cebeci T (2004) Turbulence models and their applications. Horizons Publishing Inc., Springer
Chaiko MA, Brinckman KW (2002) Models for analysis of water hammer in piping with entrapped Air. J Fluids
Eng ASME 124:194–204
Chaudhry MH (2014) Applied Hydraulic Transients. 3rd Edition, Springer
M. Besharat et al.

Dostal V, Železný V, Zacha P (2008) SIMULATIONS OF TWO-PHASE FLOW IN FLUENT. ANSYS


konference; 16th ANSYS FEM Users’ Meeting & 14th ANSYS CFD Users’ Meeting
Falk K, Gudmundsson JS (2000) Water Hammer in High-Pressure Multi-Phase Flow.Proceeding of 8th
International Conference on Pressure Surges 41–54
Ghidaoui MS, Zhao M, Mclnnis DA, Axworthy DH (2005) A review of water hammer theory and practice. App
Mec Rev ASME 58:49–76
Jaeger C (1933) Theorie generale du coup de belier. Dunod, Paris
Kader B (1993) Temperature and concentration profiles in fully turbulent boundary layers. Int J Heat Mass Transf
24(9):1541–1544
Kwon HJ (2007) Analysis of transient flow in a piping system. KSCE J Civ Eng 11:209–214
Launder BE, Spalding DB (1972) Lectures in mathematical models of turbulence. Academic, London, England
Lee TS (1994) Numerical modelling and computation of fluid pressure transients with Air entrainment in
pumping installations. Int J Numer Methods Fluids 19(2):89–103
Martins SC (2013) Dynamic of Hydraulic System Pressurization with Entrapped Air. Dissertation, University of
Lisbon, Instituto Superior Técnico, Lisbon (In Portuguese)
Martins SC, Martins NMC, Ramos HM, Almeida AB (2012) Liquid flow and entrapped behaviours in an
experimental set-up using CFD analysis, 11th International Conference Pressure Surges, Lisbon, October
24–26, 505–515
Martins NMC, Carriço NJG, Ramos HM, Covas DIC (2014) Velocity-distribution pressurized pipe flow using
CFD: accuracy and mesh analysis. Comput Fluids 105:218–230
Mathur A, He S (2013) Performance and implementation of the launder–Sharma Low-Reynolds number
turbulence model. Comput Fluids 79:134–139
Parmakian J (1955) Water Hammer Analysis. Prentice-Hall, Inc., Englewood Cliffs, NJ (Dover Reprint 1963)
Ramos HM, Covas D, Borga A, Loureiro D (2004) Surge damping analysis in pipe systems: modeling and
experiments. J Hydraul Res 42(4):413–425
Rich G (1944) Water Hammer Analysis by the Laplace-Mellin Transformations. Trans ASME 1944–45
Rich G (1951) Hydraulic transients. McGraw-Hill Book Co New York, New York
Shih TH, Zhu J, Lumley JL (1994) A new Reynolds stress algebraic equation model. NASA TM
Shih TH, Liou WW, Shabbir A, Yang Z, Zhu J (1995) A New k-ε eddy-viscosity model for high Reynolds
number turbulent flows-model development and validation. Comput Fluids 24(3):227–238
Streeter VL, Lai C (1963) Water hammer analysis including fluid friction. J Hydraul Div ASCE 128:1491–1524
Streeter VL, Wylie EB (1967) Hydraulic Transients. McGraw-Hill Book Co NY
Wilcox DC (2006) Turbulence Modeling for CFD. DCW Industries. 3rd edition
Wood FM (1937) The application of Heaviside’s operational calculus to the solution of problems in water
hammer. Trans ASME 59:707–713
Wörner M (2003) A Compact Introduction to the Numerical Modeling of Multiphase Flows. Forschungszentrum
Karlsruhe, Wissenschaftliche Berichte FZKA 6932
Wylie EB, Streeter VL (1993) Fluid transients in systems. Prentice-Hall, NJ
Zhang T, Zhou H, Wang S (2013) An adjustment to the standard temperature wall function for CFD modeling of
indoor convective heat transfer. Build Environ 68:159–169
Zhou L, Liu D, Ou C (2011) Simulation of Flow Transients in a Water Filling Pipe Containing Entrapped Air
Pocket with VOF Model. Engineering Applications of Computational Fluid Mechanics 5(1): 127–1

Potrebbero piacerti anche