Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Journal of Natural Gas Science and Engineering 92 (2021) 103989

Contents lists available at ScienceDirect

Journal of Natural Gas Science and Engineering


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

Models for predicting the test medium temperature during hydrotesting


pipe segments based on measured ambient and pipe wall temperatures
K.K. Botros a, *, J. Crowe a, V. Liu b, J. Lu b
a
NOVA Chemicals Corporation, Centre for Applied Research, 2928 16th Street NE, Calgary, Alberta, Canada
b
TC Energy, 450 1st St SW, Calgary, Alberta, Canada

A R T I C L E I N F O A B S T R A C T

Keywords: Strength and leak tests of newly constructed or modified pipeline sections or piping assemblies are required by
Gas pipeline code. This is often conducted by pressurizing incompressible medium, such as water, water-methanol or water-
Hydrotesting propylene glycol (PG) mixtures while the system is sealed. Major challenges are invariably encountered with leak
Leak test
tests, which rely on correlating changes in the test pressure to the test fluid temperature (DP/DT) to discern if a
Fluid bulk modulus
Thermal expansion
leak exists. Often, the temperature variations are never correlated to pressure variations for several reasons: 1)
Strained access to the test fluid temperature is not available for several reasons outlined in the paper, therefore, the
Un-restrained pipe measured external pipe wall temperature is taken instead and is assumed to be equal to the test fluid inside the
pipe, 2) the pipe wall temperature variations are generally significant due to variation in the ambient temper­
ature or wind speed (in the case of exposed pipe), 3) the pipe section may not be restrained from axial movement,
4) tables and calculations of DP/DT are readily available for pure water and often erroneously assumed to be
applicable to other test media such as mixtures of water-methanol or water-PG, which are vastly different than
pure water, and 5) some pipe sections may be partially exposed to ambient and partially buried. The present
work addresses these factors via development of high-fidelity models based on governing equations that accounts
for the respective effects in a more fundamental manner. It was found that it is paramount to use the correct test
medium isothermal bulk modulus and its coefficient of volumetric thermal expansion at the test conditions as
these two parameters have the most significant influence on DP/DT. Due to these properties, water-PG was
shown to result in the highest DP/DT, which poses the greatest challenges during hydrotesting in the field.
Additionally, it was found that the difference between the pipe external wall temperature and the test fluid
average temperature for the case of exposed pipe increases as Biot number increases. The developed thermal
transient models were compared to three field hydrotests of different pipe sizes, namely DN900, DN150 and
DN50, all above ground subject to cross wind and variations in ambient air temperatures.

2885; https://www.saiglobal.com, 2006), and NOM 007 ASEA 2016 (in


Mexico) (An International, 2016).
1. Introduction Often during hydrotesting, the test medium is subjected to temper­
ature variations due to many environmental factors and site conditions.
Following the completion of construction, a pipeline segment or In this case, it is recommended that consideration be given to estab­
piping assembly is required to be hydrotested to a specified test pressure lishing the pressure-temperature relationship so that the effects of such
depending on the design pressure, applicable codes and Class location. variations can be taken into account in determining the pressure
Several organizations and governmental bodies have developed stan­ gradient expected during the test period. During the temperature sta­
dards and guidelines relating to pipeline construction and hydrotesting. bilization period, however, i.e., between the time of filling and the
The most prominent regulations and standards are: ASME B31.3 (pro­ commencing of the pressure test, it is not required that a pressure-
cess piping) (ASME B31.3, 2018), B31.4 (liquid pipelines) (ASME B31.4, temperature gradient be established because pressure-temperature
2019), B31.8 (gas pipelines) (ASME B31.8, 2018), CFR 49 192/195 extrapolation is not necessarily reliable. Whereas, during the hydrotest
(DOE - 49 CFR Ch; DOE- 49 CFR Ch), CSA Z662 (in Canada) (CSA test it is utmost important to correlate changes in the test pressure to the
Z662:19), AS2885.5/AS3788 (in Australia) (https://shop.standards.go,

* Corresponding author.
E-mail address: kamal.botros@novachem.com (K.K. Botros).

https://doi.org/10.1016/j.jngse.2021.103989
Received 25 January 2021; Received in revised form 24 April 2021; Accepted 27 April 2021
Available online 1 May 2021
1875-5100/© 2021 Elsevier B.V. All rights reserved.
K.K. Botros et al. Journal of Natural Gas Science and Engineering 92 (2021) 103989

Nomenclature V Fluid volume (m3)


WI Direct Normal Irradiance (DNI) from the sun (W/m2)
A Pipe external surface area (m2) W.T. Wall thickness (m)
Bi Biot number (− ) z Axial coordinate along the pipe (m)
BT Fluid isothermal bulk modulus (Pa) α Pipe material linear thermal expansion coefficient (1/oC)
D Pipe external diameter (m) αi Thermal diffusivity of substance (i) (m2/s)
DP Change in pressure (Pa) φ Azimuthal coordinate (radians)
DT Change in temperature (oC) θ Angle between solar radiation and pipe surface (radians)
E Young’s modulus of the pipe material (Pa) ρ Fluid density (kg/m3)
H Depth of burial from tope of the pipe to ground surface (m) τ Normalized time (− )
k Thermal conductivity (W/m.oC) ν Poisson ratio of pipe material (− )
m Fluid mass inside the pipe segment (kg) ψ Surface absorptivity of the external pipe wall (− )
P Fluid pressure (Pa)
r Radial coordinate (m) Subscript
R Internal pipe radius (m) a ambient air
Ro Outer pipe radius (m) i initial condition at time zero
t Pipe wall thickness (in Eq. 3 through 9) (m) m test medium
t Time (Eq. (10) onward) (s) p pipe
T Temperature (oC) s soil
Uo External heat transfer coefficient (W/m2.oC) w water

test medium temperature to discern if a leak exists. In practice, the validated against hydrostatic field data on a DN900, DN150 and DN50
temperature variations are never correlated to pressure variations for above ground pipe assemblies at compressor stations at different sites.
several reasons: 1) the temperature is measured on the outer wall of the
pipe which is assumed to be the same as the test fluid inside the pipe, 2) 2. Literature review
the pipe wall temperature variations are generally significant due to
variation in the ambient temperature or wind speed (in the case of In order to apply the fundamental pressure-temperature relationship
exposed pipe), 3) the pipe section may be not restrained from axial during hydrotesting, the thermodynamic properties of the test medium
movement, 4) the properties of the test fluid if it is a mixture of water- (isothermal bulk modulus and coefficient of volumetric thermal
methanol or water-glycol at different mix ratios are vastly different expansion) needed to be quantified as function of pressure and tem­
from pure water, and 5) some pipe sections may be partially exposed to perature. The thermodynamic properties of water are well known
ambient and other portion is buried. Invariably, these factors leave the (Bahadori and Vuthaluru, 2009), however the thermodynamic proper­
project team unable to decide on the spot as to whether to accept or ties of water/methanol, water/ethylene-glycol, water/propylene-glycol
reject the hydrotest results. mixtures are less understood. The properties of the water-methanol
The present work is aimed at addressing the above five challenges mixture at different temperatures and pressures have been studied by
facing a field engineer on the ground attempting to determine if there is several authors (Soetens and Bopp, 2015; Kume et al., 2005; Huc et al.,
a leak or not, and to accept or reject hydrotest results. In particular, the 2015; Kubota and Tsuda, 1979; Hartono and Kim, 2004). Soetens and
first challenge is related to the measurements of the test fluid temper­ Bopp (2015) simulated the mixing properties of water-methanol mix­
ature. For pipeline sections, often there are no thermowell connections tures over the entire range of mole fractions and found good agreement
available as per design, especially for buried pipelines. While thermo­ with experimental results. Kume et al. (2005) developed an equation of
wells could be added on the “test-head”, they do not represent the state for methanol at temperatures of − 98 ◦ C–300 ◦ C, and pressures up
temperature of the test fluid inside the full stretch of the pipeline. For to 200 MPa. Huc et al. (2015) used Monte Carlo simulations to develop
small size facility piping, there may not be thermowells available either, an isotropic model for water/methanol mixtures and found that it
or while it is possible to install one on the so-called “test-tree”, it would correctly described the density dependence of water-methanol on tem­
be at the very end of the pipe with the “jumper-line” separating it from perature and composition. Kubota and Tsuda (1979) studied the specific
the piping under hydrotesting. For large bore facility piping, there might volume and viscosity of methanol/water mixtures in relation to tem­
be thermowells available, however they typically have very short perature, pressure, and composition. When using water/methanol
insertion lengths, which in reality, they measure more or less the test mixtures, challenges arise due to the coefficient of thermal expansion of
fluid temperature near the pipe wall and not the bulk. methanol being 7–9 times higher than that of water, and therefore much
The paper provides a methodology to address the influence of the higher changes in pressure occur with changes in temperature, i.e.,
principal properties of the test medium combined with the relevant DP/DT (Hartono and Kim, 2004). This makes it difficult to detect leaks
geometrical parameters of the pipe segment and environment conditions with water/methanol mixtures since the pressure change due to loss of
during the test, via development of high-fidelity thermal transient fluid from a leak cannot be easily resolved.
models based on governing equations that accounts for the respective Propylene Glycol (PG) can also be used instead of methanol due to its
effects in a more fundamental manner. This is described in detail in lower environmental impact. The literature contains properties of
Section 3. Section 4 describes two-dimensional thermal transient (un­ water-PG mixture only for (50%wt) (Sagdeev et al., 2017; Zhuravlev,
steady) models developed to determine the relationship between tem­ 1992; Guignon et al., 2010). From this data, the present work developed
poral variations in ambient temperature and the measured pipe outer an empirical correlation for the density of water-PG (50%wt) mixture as
wall temperature and the resulting temporal variations in the test fluid follows:
mean temperature: one for buried pipes and another for exposed pipes. ( ) ( )
Effects of variations in ambient temperatures are presented in Section 5
ρwater− PG(50%wt) = − 0.0019T 2 +0.5262T +1050.4 + − 0.0002P2 +0.3374P
in a normalized manner for different pipe geometries, test fluid prop­ (1)
erties, and environmental conditions. The developed models were

2
K.K. Botros et al. Journal of Natural Gas Science and Engineering 92 (2021) 103989

where: T in K, and P in MPa.


Fig. 1 shows an example of density vs. temperature at P = 12 MPa of
pure water, water-methanol (50%wt) and water-PG (50%wt) mixtures.
The latter was based on the developed correlation (1) above, while the
formers are based on REFPROP (Kunz and Wagner, 2012; Lemmon et al.,
2010). Note that the density of water-PG (50%wt) is highest over the full
range of temperature. However, it is not the absolute value of density
that determines relationship DP/DT, but rather the thermodynamic
derivatives. The resulting derivative thermodynamic properties
(dρ/dT)P and isothermal bulk modulus for three test fluids are also
compared in Figs. 2 and 3, respectively.
The California State Lands Commission procedure for hydrotesting
recommends that aboveground pipe tests be conducted on overcast and
cool days when possible to minimize thermal heating of the test media.
This reduces the amount of bleeding required so as not to overpressure,
which may be significant for larger sections of above ground piping on
Fig. 2. Comparison of (dρ/dT)P of pure water, water-methanol (50%wt) and
sunny/hot days (Procedure for the Hydro, 2003). Types of ambient
water-PG (50%wt) at different temperatures and 12 MPa.
conditions that should be avoided are hot, extremely windy and rainy
conditions. Multiple temperature recorders should be used when a sig­
nificant difference of the pipe fluid temperature is expected, such as test
sections having both above ground and below ground pipes, above
ground sections which vary in pipe diameters, and above ground seg­
ments exposed to sun and shade. Khatami noted that since temperature
may differ from one side of the pipe to the other, the surface temperature
of the pipe should be measured at 0, 90 and 180 degrees of the cross
section, and the average value of these should be used as the pipe
temperature at the time of testing (Khatami et al., 2020). The California
State Lands Commission states that for aboveground sections, the tem­
perature probe should be attached underneath the pipe and thermally
insulated from the ambient. It is recommended that insulation extend
one pipe diameter on each side of the probe (Kirkwood and Cosham,
2000). There is, however, some debate when it comes to insulating a
segment of the pipe, the concern being, an insulated location may not be
representative of the whole pipe segment. Insulation will alter the heat
transfer between the outer surface of the pipe and ambient conditions,
and this local point may not provide an accurate representation of the Fig. 3. Comparison of isothermal bulk modulus of pure water, water-methanol
whole pipe segment. Further investigation on this should be conducted (50%wt) and water-PG (50%wt) at different temperatures and 12 MPa.
in future work.
Entrapped air will also affect the pressure-temperature calculations, test section as it displaces a significant volume of compressible air
since the expansion coefficient of air is not the same as water or other (Matta, 2017).
incompressible test medium, and proper measures should be taken to Matta examined the collective effects of leakage, temperature
avoid air entrapment (Matta, 2017). A small amount of entrapped air changes and entrapped air during hydrostatic testing (Matta, 2017). He
can significantly affect the compressibility of the test fluid, and as a developed equations to quantify the effects of temperature changes and
result if a leak is present, the pressure loss may be less than expected leakage rates on changes in pressure, as well as the effect of trapped air
(Kunz and Wagner, 2012). The first indication of large amounts of in the pipe on the test pressure. The results of the work indicate that
entrapped air usually occurs during initial pressurization, where the small leaks are most easily detectable in short test sections with stable
pressure increase would not be immediate as water is injected into the temperatures and little air entrapment. The magnitude of the pressure
changes, resulting from the temperature variation and leaks throughout
the test, is highly dependent on the starting temperature and pressure.
He found that the influence of temperature variations on the test pres­
sure are minimized at colder temperatures, water at 4 ◦ C, and increase
with increasing test temperature (Matta, 2017).
Concluding from the above cited literature, it is clear that an issue
exists with regard to what constitutes the actual and correct test medium
temperature and the commonly made assumption that it is equal to the
external wall temperature of the pipe as discussed earlier. Hence, the
external wall temperature is measured and taken as a ‘proxy’ of the test
medium temperature. As such, procedures are developed by individual
companies, e.g. (Procedure for the Hydro, 2003) to provide best prac­
tices to install and shield these external temperature sensors from sur­
rounding environment, location on the pipe wall in the azimuthal
direction and along the stretch of the pipe segment, as well as the limited
number of sensors from which an average value is taken. In some cases,
only one temperature sensor is available, and its location is not the
Fig. 1. Comparison of density of pure water, water-methanol (50%wt) and optimum either. Despite all these precautions and specifications, still
water-PG (50%wt) at different temperatures and 12 MPa. remains the challenge that the pipe external wall temperature is

3
K.K. Botros et al. Journal of Natural Gas Science and Engineering 92 (2021) 103989

invariably NOT equal to the test medium temperature. This is the [ ( )]


D( ) 1
principal motivation of the present work, in an attempt to develop 0=ρ V 1 − ν2 + δP
means to discern the test medium temperature from external wall tem­ tE BT
[ ( ) ]
perature measurements, properties of the test fluid and geometry and + ρ V 2(1 + ν)α +
1 ∂ρ
δT (for Restrained Pipe) (6)
properties of the pipe segment under hydrotest. ρ ∂T P
Similarly:
3. Governing equations [ ( ) ( )]
D 5 1
0=ρ V − ν + δP
A pipeline segment of piping assembly containing a fixed mass of tE 4 BT
[ ( ) ]
hydrotest medium can be treated as a closed but variable volume system, 1 ∂ρ
+ ρ V 3α + δT (for Un − Restrained Pipe) (7)
while small amount of this medium can bleed out (leak) through a ρ ∂T P
pinhole. Heat can be transferred across the boundaries, allowing the
temperature of the medium to change (increase or decrease). Because Finally:
[ ( ) ]
the temperature can change, the pressure and density of the medium can
also change. The pipe walls are flexible and subject to thermal expan­ ( ) ( ) − 2(1 + ν)α + 1ρ ∂∂ρT
δP DP
sion, so the volume of the system will change as a result. Additionally, an = = [ ( )]P (for Restrained Pipe) (8)
δT DT D
increase in internal pressure causes a section of pipe to stretch, which tE
(1 − ν ) + B1T
2

increases its volume.


[
( ) ]
Generally, an increase in temperature will cause the pipe to expand
radially and axially and increase in volume. How much the pipe expands ( ) ( )3α + ρ1 ∂∂ρT

δP DP
depends on the restraint of the pipe. For a short section of pipe above = =[ ( ) ( P)] (for Un − Restrained Pipe) (9)
δT DT D 5
ground, for example, the ends of the pipe may be capped, and the in­ tE 4
− ν + B1T
ternal pressure will act upon the close ends. In this case, the pipe is
acting as a cylindrical pressure vessel, where the expansion occurs in Unlike pneumatic testing, the sensitivity of changes in δP with
two directions – radially and axially. This condition is referred to as respect to changes in δT (which will be denoted DP/DT from here on),
“unrestrained”. For a long section of buried pipeline, however, the due to pipe restrain/un-restrained conditions, thermal expansion and
friction of the soil prevents the pipe from expanding axially. This is the pipe compliance is significant and cannot be neglected. Furthermore,
“restrained” case, and the pipe is assumed to expand only in the radial two other primary properties of the test fluid come to play an important
direction. Any resistance to radial expansion of the pipe by the soil is role in the relationship between DP and DT according to Eqs. (8) and (9),
neglected. This is also true for above ground pipe segments when the namely, the isobaric coefficient of thermal expansion (∂ρ/∂T)P and the
two ends are rigidly anchored and restrained from axial movements. isothermal bulk modulus of the test fluid, BT . Both parameters are de­
The fundamental governing equations of the effect of pressure and rivatives of the functional density with respect to temperature and
temperature on the volume of the pipe for the restrained and unre­ pressure, respectively, as was shown in the previous Section 2.
strained cases without leak can be expressed as follows: A numerical example based on Eqs. (8) and (9), is provided to
demonstrate the role and effects of the aforementioned parameters on
m=ρ V (2) the relationship DP/DT. Consider a DN400 pipe segment of wall thick­
ness = 7.925 mm (hence internal diameter, I.D. = 390.55 mm), E = 207
δm = ρ δV + V δρ (2a) GPa, ν = 0.28 and the coefficient of linear thermal expansion of the pipe
material, α = 1.17 × 10− 5 m/ C. Based on the principal properties the

where, in the case of no leak or no test fluid addition or bleeding off,


three test fluids, water, water-methanol (50%wt) and water-PG (50%
δm = 0.
wt), DP/DT can be calculated for this pipe segment, and the results are
First term in (2a):
shown in Fig. 4.
[ ]
D( ) It is shown that water-PG (50%wt) exhibits the highest stiffness
ρ δV = ρ V 1 − ν2 δP + 2(1 + ν)α δT (for Restrained Pipe) (3)
tE compared to water-methanol (50%wt) and pure water. For instance, at
T = 10 ◦ C and P = 12 MPa, DP/DT is equal to approx. 120 kPa/ C for

[ ( ) ]
pure water, 780 kPa/ C for water-methanol (50%wt), and 850 kPa/ C
◦ ◦
D 5
ρ δV = ρ V − ν δP + 3α δT (for Un − Restrained Pipe) (4)
tE 4 for water-PG (50%wt). This indicates that water-PG (50%wt) is the
The difference in the formulation of Eq. (3) vs. Eq. (4) is due to the
boundary condition effects on the relationship between strain vs. stress
in the three-principal directions (axial, radial and azimuthal), the deri­
vation of which can be found in any solid mechanics textbook.
Second term in (2a):
[( ) ( ) ]
∂ρ ∂ρ
V δρ = V δP + δT
∂P T ∂T P
[( ) ( ) ] (5)
ρ ∂ρ
=V δP + δT
BT ∂T P

Substituting in (2a) and rearranging:


[ ] [( ) ( ) ]
D( ) ρ ∂ρ
0=ρ V 1 − ν2 δP + 2(1 + ν)α δT + V δP + δT
tE BT ∂T P
or;

Fig. 4. Comparison of DP/DT of a DN400 pipe segment based on different test


fluids at 12 MPa.

4
K.K. Botros et al. Journal of Natural Gas Science and Engineering 92 (2021) 103989

stiffest fluid followed by water-methanol (50%wt), while pure water is [ ( ) ( )]


∂Tp 1 ∂ ∂Tp 1 ∂2 Tp
the least stiff. Pipe wall: = αp r. + 2 (11)
∂t r ∂r ∂r r ∂φ2
The term “stiffness” here is used rather loosely to indicate the relative
value of the change in DP with respect to a change in DT, i.e., DP/DT. [ ( ) ( )]
∂Ts 1 ∂ ∂Ts 1 ∂2 Ts
This illustrates the challenges in a hydrotest if water-PG or water- Soil: = αs r. + 2 (12)
∂t r ∂r ∂r r ∂φ2
methanol mixtures are used instead of pure water as fluctuation of DP
will be relatively high to the extent that it could hinder the ability to
where, α represents thermal diffusivity, r, radial coordinate, T, tem­
discern a pinhole leak. Unfortunately, hydrotesting in sub-freezing
perature and t, time. Subscripts (m, p, s, a) denote test medium, pipe
ambient temperature prohibits the use of pure water and mixing with
material, ground soil, and ambient air, respectively. Solutions of the
methanol or PG is necessary. Depending on how low the ambient tem­
above transient thermal governing equations in the form of spatial-
perature gets, it is important to select the right concentration of either
temporal temperature distribution can be expressed as function of
methanol or PG to prevent freezing, but not to over-mix unnecessary to
dimensionless parameters as follows:
avoid the issue of undesirable high DP/DT. Another observation in Fig. 4
( )
is that the difference in DP/DT between restrained vs. unrestrained is T − Ta αs αp r t αm
T= =f ; ; ; τ= 2 (13)
small, and would progressively become smaller for smaller pipe sizes, Tmi − Ta αm αm R Ro
and the opposite occurs as the pipe size gets larger.
where, τ, is the dimensionless time and subject to boundary conditions.
4. Pipe wall vs. test medium temperatures Equation (13) describes the thermal field in 0 > r > R close to the pipe,
which indicates that the relationship between test medium temperature
The formulation of the governing equations and analysis presented in and the pipe external wall temperature is driven primarily by soil-to-
the previous Section assumes that the measured outer wall temperature medium thermal diffusivity ratio (αs/αm), since the ratio (αp/αm) is
of the pipe section under hydrotesting is equal to the test medium essentially constant. Furthermore, the solution is considered self-similar
temperature. It is well known from numerous field measurements that with respect to r/R, and that the time scale is proportional to square of
this is not the case, albeit that the test medium temperatures are rarely the pipe radius, R. Note that, the effect of the burial depth, H, is
measured (ASME B31.3, 2018; ASME B31.4, 2019; ASME B31.8, 2018; inconsequential as it is almost always the same for most pipe sizes
DOE - 49 CFR Ch; DOE- 49 CFR Ch; CSA Z662:19; Procedure for the (approx. 1.0 m). Fig. 5 shows details of the simulation space (left), and
Hydro, 2003; Khatami et al., 2020). To validate this assumption, a model grid structure (center). The boundary condition at the bottom and
transient thermal analysis was performed using ANSYS software right of the domain shown as Ts is constant and is equal to Ta at the
(ANSYS, 2020) to discern the relationship between the pipe external ground surface.
wall temperature and the test medium average temperature. This Sec­ The FEA-based simulations were performed in ANSYS Mechanical
tion describes the basis of two models: one for buried pipes and the other 2020 R1 using the transient thermal solver. The solid domain was
for above ground pipes. Results of these models are presented in Section meshed using PLANE77 elements, a higher order 8-node thermal
5. element used for 2-D steady-state or transient analysis. The transient
model was solved using a distributed sparse matrix direct solver. A mesh
4.1. Buried pipe independence study was conducted for a DN400 pipe to determine if the
resulting water and pipe temperatures reached a mesh independent
Fig. 5 depicts ANSYS model for the buried pipe along with salient solution. When the mesh size was decreased from 2 mm to 1 mm (8.5k
model parameters for the transient model. Ignoring the derivative in the elements to 33k elements) the resulting average water temperature
axial direction (z coordinate), the thermal transient governing equations changed by only 0.02 ◦ C. Therefore, a mesh size of the order of 1 mm
can be expressed in two-dimensional domain along the radial direction, was adopted for DN400 pipes, and scaled up/down for other pipe sizes.
r, and the azimuthal direction, φ, as follows: Example results of the model predictions are shown in Fig. 6 for the
[ ( ) ( )] following pipe geometry, soil characteristics, initial conditions, and the
∂Tm 1 ∂ ∂Tm 1 ∂2 Tm thermal diffusivity ratio (αs/αw = 2.8) at different time stamps of 1, 3, 6
Test medium: = αm r. + 2 (10)
∂t r ∂r ∂r r ∂φ2

Fig. 5. Schematic of ANSYS model for buried pipe to determine the relationship between the external pipe wall temperature and the test medium average tem­
perature inside the pipe.

5
K.K. Botros et al. Journal of Natural Gas Science and Engineering 92 (2021) 103989

Fig. 6. Example results of the temperature field inside and outside a buried pipe at different time for diffusivity ratio (αs/αw = 2.8).

and 9 h from the start of the hydrotest. The following parameters are air in the presence of cross wind as follows:
used in this example, assuming that the test medium is pure water: ∫ ( ) ∫
∂T ( )
kp . .dA = AUo Tp − Ta − ψ .WI .cos(θ).dA (14)
∂r r=Ro
• Initial conditions: Tmi = 20 ◦ C, Tpi = 20 ◦ C, Tsi = 10 ◦ C. A A
• Constant parameters: Ta = 10 ◦ C, αw = 0.143 × 10− 6 m2/s, αp = 12 ×
10− 6 m2/s, αs = 0.4 × 10− 6 m2/s, R = 0.3937 m (DN800), W.T. = where, Ro is the outer pipe radius, Uo is the external heat transfer co­
0.0127 m, H = 1.0 m efficient which is determined from an applicable Nusselt number cor­
• Time steps: Adaptive time steps; min. = 0.001s (during initial stage relation with cross flow Reynold number (based on cross wind speed,
when temperature changes are fast), max. = 500s (after first 1–2 h outer pipe diameter, and ambient air density and viscosity) and Prandtl
when temperature changes are slow). number of ambient air (Kirkwood and Cosham, 2000). The second term
on the R.H.S. of Eq. (14) is related to heat absorption due to solar ra­
Here, the ambient temperature is assumed to remain constant at diation, where ψ is the surface absorptivity of the external pipe wall, WI
10 ◦ C during the entire duration of the hydrotest, while the test medium is the Direct Normal Irradiance (DNI) from the sun during the time of the
(water) and pipe wall temperatures were assumed to be at 20 ◦ C at time test at the hydrotest site location, and the integral is over the part of the
= 0. This assumption is revisited and discussed in Section 5. Fig. 7 shows external surface area facing the sun (Holman, 2009). If solar radiation is
the result of the simulation in terms of normalized external pipe wall not considered, then Eqs. (10) and (11) will not have the azimuthal
temperature and cross-sectional area averaged test fluid temperature terms on the respective R.H.S. Again, the axial direction is not consid­
inside the pipe vs. normalized time, τ, for different (αs/αw) ratios. ered here assuming no temperature gradient in the z-direction.
Likewise, solution of the pertinent Eqs. (10), (11) and (14) in the
form of spatial-temporal temperature distribution can be expressed as
4.2. Above ground pipe
function of dimensionless parameters as follows:
( )
The thermal transient governing equations for exposed (above T − Ta Uo Ro αp r t αm
ground) pipe segments are similar to the buried pipe in so far as the test T= =f ; ; ; τ= 2 (15)
Tmi − Ta km αm R Ro
medium (Eq. (10)) and pipe wall (Eq. (11)) again ignoring the derivative
in the axial direction (z coordinate), but the soil equation (Eq. (12)), is Note that the first term on functional R.H.S. is Biot number defined in
replaced with the heat transfer between the pipe outer wall and ambient the present work as:
Uo R
Bi = (16)
km

where, km is the thermal conductivity of the test medium. Likewise, the


relationship between test fluid temperature and pipe external wall
temperature is driven solely by Bi number, since again the ratio (αp/αm)
is essentially constant. Similarly, the solution is considered self-similar
with respect to r/R, and that the time scale is proportional to square
of pipe radius, R. Fig. 8 shows details of the simulation space and grid
resolution adopted in ANSYS thermal transient simulations performed in
the present work.
Example results for the conditions and pipe geometry of Fig. 8 are
shown in Fig. 9 for Bi number = 2 at different time stamps of 1, 3, 6 and
9 h from the start of the hydrotest. Note that ambient temperature is
assumed to remain constant at 10 ◦ C in this example as well, an
assumption which will be addressed in the next Section. Fig. 10 shows
Fig. 7. Normalized external pipe wall temperature and area integrated test the result of the simulation in terms of normalized external pipe wall
fluid temperature inside the pipe vs. normalized time, τ, for different (αs/αw) temperature and area integrated test fluid temperature inside the pipe
ratios for buried pipe segments.

6
K.K. Botros et al. Journal of Natural Gas Science and Engineering 92 (2021) 103989

and surrounding temperatures, all reached the same temperature before


pressurizing and commencing of the hydro tests. Ambient temperature
was then assumed to be ramping up at the rate Ṫa . The intent was to
determine how the pipe wall temperature and test fluid temperature
track this ambient temperature. The results from ANSYS simulations are
normalized as follows:
( )/( )
Ṫ a − Ṫ w Ṫ a − Ṫ p (17)

where the time derivatives of the various temperatures are defines as


follows:

Ṫ a = (dTa /dt)
Ṫ w = ((dTw//dt)) (18)
Ṫ p = dTp dt

5.1. Buried pipe

Fig. 11 shows ANSYS model parameters of the same DN800 buried


pipe segment used in previous examples, and results in terms of pipe
Fig. 8. Schematic of ANSYS model for above ground pipe to determine the
wall temperature and average water temperature over a period of 10 h
relationship between the external pipe wall temperature and the test medium from the start of the hydrotest. In this example, after sufficiently long
average temperature inside the pipe. stabilization time, the test water temperature, pipe wall temperature
and ambient temperature all assumed to have reached equilibrium at
vs. normalized time, τ, for different Bi number. 10 ◦ C. At the start of the hydrotest, the ambient temperature is assumed
to be increasing at a rate of 1 C/h. The results show that the test water

temperature has hardly increased, and the pipe wall temperature


5. Effects of variations in ambient temeperture
increased only by 0.04 ◦ C, which is rather negligible. Notice the tem­
perature contours in the ground soil after 10 h in Fig. 11 (bottom left),
The analysis conducted thus far assumed constant temperature of
that it hardly penetrated to the depth at the top the pipe.
ambient air, and that the hydrotest test procedure commences imme­
Similar simulation was conducted on DN50 pipe segment. Fig. 12
diately after filling where there could be a vast difference between
ambient air temperature and the test medium temperature (e.g., 10 ◦ C
and 20 ◦ C, in the two examples shown in Figs. 6 and 9). Clearly, if there
is a variation in ambient temperature during the test, it will affect the
relationship between the pipe wall temperature and the average me­
dium temperature over the duration of the test.
In practice, the actual hydrotest procedure does not commence
immediately after filling, either purposely or due to time taken to check
for entrained air, sequence of multiple bleed-off and injection of test
medium to replace the bled fluid/air mixture, etc. Following this step,
the field crew usually leave the system bottled-in for extended period
(typically one day, or even longer) purposely to stabilize the system and
allow for further entrained air to migrate to a high elevation point for
subsequent bleed-off/re-injection procedure afterward. Often, the test
medium temperature, pipe wall temperature, soil and ambient air
temperature come very close to equalizing before commencement the
Fig. 10. Normalized external pipe wall temperature and area integrated test
hydrotest.
fluid temperature inside the pipe vs. normalized time, τ, for different Bi number
Simulation cases were conducted in ANSYS following a full stabili­
for above ground pipe segments.
zation period such that the test fluid temperature, pipe wall temperature

Fig. 9. Example results of the thermal field inside an above ground pipe at different time for Bi = 2.

7
K.K. Botros et al. Journal of Natural Gas Science and Engineering 92 (2021) 103989

Fig. 11. ANSYS model parameters and results of buried DN800 pipe segment subject to hydrotesting while ambient temperature is changing at a rate of 1 C/h after

full stabilization temperature of 10 ◦ C.

shows ANSYS model parameters and results of pipe wall temperature reason is that the pipe is now exposed to the ambient with a thermal
and average water temperature over a period of 10 h from the start of the resistance corresponding to external heat transfer coefficient (Uo), via
hydrotest. The results show that the test water temperature increased the dimensionless Bi number, as opposed to the case of buried pipes
only by 0.024 ◦ C, and likewise, the pipe wall temperature increased by where the soil (massive mass) hinders the penetration of ambient tem­
only 0.033 ◦ C, both of which are rather negligible. perature variations to reach the pipe wall and affect the water temper­
ature inside the pipe. In the present case, the pipe wall temperature
increase depends on the Bi number as shown in Fig. 13. It is clear that as
5.2. Above ground pipe Bi number increases, the gap (or difference) between the pipe wall
temperature and the area average water temperature widens, which lead
The same investigations were carried out for above ground pipe to higher dimensionless temperature rates defined by Eq. (17).
segments. In this case, we considered one pipe size of DN400 at varying It should be noted that based on the dimensionless parameters of Eq.
Bi numbers. As will be seen later, the results are independent of the pipe (15), the normalized rate of change in temperatures defined is Eq. (17) is
size as long as the time is normalized by the dimensionless parameter, τ. solely dependent on Bi number up to a given dimensionless time, τ,
Fig. 13 shows results in terms of pipe wall temperature and average irrespective of the pipe size. Therefore, based on the results of Fig. 13,
water temperature over a dimensionless time up to τ = 0.14 (corre­ values of (Ṫa − Ṫw )/(Ṫ a − Ṫ p ) are plotted against Bi number in Fig. 14,
sponding to 10 h for the pipe size of DN400 in this example) from the which can be fitted in the form of Eq. (19). It is important to note that
start of the hydrotest. Again, time 0 starts after long stabilization time,
this equation is applicable only up to τ = 0.14. Further work is underway
the test water temperature, pipe wall temperature and ambient tem­ to extend this relationship to higher Bi numbers, as well to include τ as
perature all assumed to have reached equilibrium at 10 ◦ C. At the start of
another dependent dimensionless parameter in the relationship to
the hydrotest, again the ambient temperature is assumed to be
determine (Ṫa − Ṫw )/(Ṫ a − Ṫp ).
increasing at a rate of 1 C/h. Other parameters used in the simulations

( )/( )
are: αw = 0.143 × 10-6 m2/s, kw = 0.6 W/m.C, αp = 12 × 10-6 m2/s, pipe
Ṫ a − Ṫ w Ṫ a − Ṫ p = 1.0 + 0.1874 Bi (19)
W.T. = 9.52 mm, and R = 0.19368 m.
The results show different trends than that of buried pipes. The

8
K.K. Botros et al. Journal of Natural Gas Science and Engineering 92 (2021) 103989

Fig. 12. ANSYS model parameters and results of buried DN50 pipe segment subject to hydrotesting while ambient temperature is changing at a rate of 1 C/h after

full stabilization temperature of 10 ◦ C.

temperatures stated above. Note that the rate of increase of water


6. Comparison and validation against field data
temperature is lagging that of the pipe wall temperature. This translates
to 0.7744 ◦ C over a period of 4 h. Following the DP/DT formulation in
In this Section, three hydrotest data sets of one large diameter pipe
Section 3, and based on average test pressure of 16835.5 kPa, and
segment and two small diameter pipe assemblies were collected over the
average temperature of 17 ◦ C, the predicted DP from Eq. (8) for
past year on TC Energy’s NGTL pipeline system in Alberta, Canada. All
restrained pipe is calculated to be = 216.5 kPa, and from Eq. (9) for un-
these pipe segments were installed above ground at different compressor
restrained pipe to be = 207 kPa, both of which are close to the measured
station sites, and hence the ambient air condition in terms of tempera­
DP of 209 kPa. No leak was observed anywhere along the pipe segment
ture variations and wind speeds were principal parameters involved in
during both strength and leak tests.
the analysis.
The second field hydrotest was conducted on above ground DN150
The first field hydrotest was conducted on an above ground DN900
Sch. 40 pipe assembly, 15 m long and includes five 90-degree elbows.
pipe of W.T. = 22.5 mm. The data of a 4-h strength test followed by 4-h
Here, the strength/leak tests were carried out simultaneously over
leak test are plotted in Fig. 15 in terms test pressure, external pipe wall
approx. 2 h. Fig. 16 shows plots of the test data in terms test pressure,
temperature (measured at 6 o’clock position) and ambient temperature.
and ambient and pipe wall temperatures. Since the pipe diameter is
During the first 4-h of the strength test, the ambient temperature is not
small, it follows from Eq. (19) and Fig. 14 that the test fluid temperature
constant but increases approximately linear at a rate of 1.9273 C/h.

would track closely to the pipe wall temperature, i.e. Ṫ w = Ṫ p , since Bi


Over the 4-h period of the strength test, the static pressure increased by
number for this pipe size is expected to be ≪ 1. In this test, the measured
209 kPa. The pipe external wall temperature also increases at an average
DP/DT is shown to be 442 kPa/0.62 ◦ C = 716 kPa/ C. Following the DP/

rate of 0.8099 C/h. The site meteorological data reported at this time

DT formulation in Section 3, the calculated DP/DT from Eq. (9) at


indication average wind speed of 2.5 km/h (0.7 m/s), therefore, the
average P = 1089 kPa, and T = 21.17 ◦ C for un-restrained pipe (due the
external heat transfer coefficient, Uo was calculated to be = 3.47 W/m2.
presence of the five elbows) was found to be = 334 kPa/ C. Here, the

K, and hence Bi = 2.64. The resulting rate of increase of the water


measured DP/DT is almost 2 times that predicted which is on the
temperature was calculated from Eq. (19) to be 0.19356 C/h, based on

opposite direction of potential leak. This was found to be due to the


Bi = 2.64, and the measured rate of change of ambient and pipe wall
possibility that water temperature is not uniform along the length of the

9
K.K. Botros et al. Journal of Natural Gas Science and Engineering 92 (2021) 103989

Fig. 13. ANSYS model results of above ground DN400 pipe segment subject to hydrotesting while ambient temperature is changing at a rate of 1 C/h after full

stabilization temperature of 10 ◦ C, Bi = 0.74, 2.9, 5.8, 8.4, 20 and 40.

kPa, and T = 9.44 ◦ C for un-restrained pipe was calculated by Eq. (9) to
be = 160 kPa/ C. The trend was opposite to the previous test, where here

the measure DP/DT is lower than predicted. No leak was visually


observed using snoop-soap tester at fittings and weld location, and
hence it was concluded that the water temperature is not uniform along
the length of the 102 m long DN50 pipe assembly. In this case, no leak
was visually detected during the test, and hence it was concluded that
(similar to the previous DN150 test) the water temperature is not uni­
form along the length of the 102 m long DN50 pipe assembly.

7. Discussion and conclusions

The following summary of findings and general conclusions can be


drawn from the present work:

Fig. 14. Normalized rate of change of temperatures vs. Bi number for above 1. The relationship between changes in test pressure due to changes in
ground pipe segments subject to hydrotesting while ambient temperature medium temperature, which is commonly denoted DP/DT is
is changing. dependent on two principal properties of the test medium, namely,
the isothermal bulk modulus and the coefficient of volumetric ther­
pipe segment, since the location where the pipe wall temperature was mal expansion of the test medium, along with four other less influ­
measured was at the very end of the pipe close to the hydrotest head ential parameters of the pipe segment itself, namely, coefficient of
with massive valves and manifold fittings. Hence the measured wall linear thermal expansion of the pipe material, Poisson ratio, Young’s
temperature was not truly representative of the entire 15 m long pipe modulus and diameter to wall thickness ratio. Restrain vs. un-
assembly. An important finding that needs future work. restrain conditions of the pipe segments under hydrotesting gener­
The third field hydrotest was conducted on above ground DN50 Sch. ally have the least effects on DP/DT.
XS pipe segment, 102 m long and includes seven 90-drgree elbows. Here, 2. By far the most important parameter that has the highest influence
a 4-h strength test followed by 4-h leak test, as shown in Fig. 17. Again, on DP/DT is the test medium isothermal bulk modulus followed by
since the pipe diameter is small, it follows that Ṫw = Ṫp , since Bi number the coefficient of volumetric thermal expansion of the test medium. It
for this pipe size is expected to be ≪ 1, according to Eq. (19), where was found that water-propylene glycol (PG) has the highest bulk
modulus compared to water-methanol, while pure water has the
limBi→0 (Ṫ a − Ṫ w )/(Ṫa − Ṫp ) = 1.0. During the strength test, the
least value. As such, testing with water-PG often poses a challenge as
measured DP/DT is shown to be 6.9 kPa/ C, while the predicted DP/DT

any small variation in the water-PG temperature would result in a


from Eq. (9) at average P = 15411.5 kPa, and T = 10 ◦ C for un-restrained
relatively high change in pressure. This poses challenges during
pipe (due the presence of the five elbows) was calculated to be = 178
hydrotesting in cold climates at sub-zero temperatures where pure
kPa/ C. Similar trend was obtained for the leak test, where the measured

water cannot be used. Use of water-PG mixtures makes it difficult to


DP/DT is shown to be 87.1 kPa/ C, while DP/DT at average P = 14432

determine whether a drop in pressure is due to a leak or due a fluid

10
K.K. Botros et al. Journal of Natural Gas Science and Engineering 92 (2021) 103989

Fig. 15. Pressure and temperature vs. time profiles during strength and leak test of DN900 above ground pipe segment (Test date: 20 July 2020).

Fig. 16. Pressure and temperature vs. time profiles during strength and leak test of DN150 Sch 40 above ground 15 m long pipe assembly (Test Date: 01
August 2020).

temperature change which following the models described in this is completed and trapped air is bled off as described in (Kirkwood
paper, can properly be determined from inference from measure­ and Cosham, 2000; Matta, 2017). Generalized charts were devolved
ments of the external pipe wall and ambient temperatures. (not presented in this paper) based on extensive ANSYS thermal
3. The most important finding of the present work is that the measured transient simulations which reveals the actual test medium temper­
external pipe wall temperatures (ASME B31.3, 2018; ASME B31.4, ature in relation to the measured external pipe wall temperature in a
2019; ASME B31.8, 2018; DOE - 49 CFR Ch; DOE- 49 CFR Ch; CSA normalized manner.
Z662:19; Procedure for the Hydro, 2003; Khatami et al., 2020) are 5. The second investigation assumed ample time for temperature sta­
generally not representative of the actual average test medium bilization to occur after filling, such that the pipe wall temperature
temperature. The latter is the temperature needed for the DP/DT and the test medium as well as the surrounding environment (soil or
evaluation to discern leak and to accept/reject test results. This has air) are all at equilibrium and assumed to have reached the same
been shown to be a critical issue during actual hydrotesting in the temperature before commencing the hydrotest procedure as recom­
field. One solution to alleviate this problem is to have a high fidelity mended in almost all standards (ASME B31.3, 2018; ASME B31.4,
thermal transient model accessible to run the hydrotest conditions 2019; ASME B31.8, 2018; DOE - 49 CFR Ch; DOE- 49 CFR Ch; CSA
overtime and discern the average test fluid temperature during the Z662:19; https://shop.standards.go, 2885; https://www.saiglobal.
hydrotest period. Another way is to measure the test fluid tempera­ com, 2006; An International, 2016). For buried pipe segments,
ture directly, but as was pointed, access to the test fluid is not regardless of pipe size, ambient temperature variation does not
available for multiple reasons discussed earlier in the paper. This is penetrate deep enough into the soil to affect either the pipe wall
issue has been investigated more rigorously in the present work, in temperature or the test medium temperature over the duration of
two ways. strength followed by leak tests (4 h each). But, for exposed pipes, the
4. The first investigation assumed no time allowed for temperature situation is different, where it was found that both pipe wall tem­
stabilization, i.e. the hydrotest commences immediately after filling perature and the test medium temperature track the ambient air

11
K.K. Botros et al. Journal of Natural Gas Science and Engineering 92 (2021) 103989

Fig. 17. Pressure and temperature vs. time profiles during strength and leak test of DN50 Sch XS above ground 102 m long pipe assembly (Test Date: 06 June 2020).

temperature, albeit at different rates depending on the external heat curation, Resources, Validation, Visualization, Writing - review & edit­
transfer coefficient, and hence the Bi number. A generalized corre­ ing. J. Lu: Conceptualization, Technical Guidance, Field Experience,
lation (Eq. (19)) has been developed in terms of a dimensionless rate Field Data Collection, Data curation, Resources, Validation, Visualiza­
of change of all three temperatures (ambient, external pipe wall and tion, Writing - review & editing.
test medium) as a function of Bi number up to certain dimensionless
time. Good agreement was demonstrated with field data of a DN900
pipe segment strength/leak test data from a recent hydrotest project. Declaration of competing interest
6. The developed models and procedures were also compared against
two small pipe diameter pipes (DN150 and DN50) recently tests for The authors declare that they have no known competing financial
strength and leak tests at different locations and at different condi­ interests or personal relationships that could have appeared to influence
tions. It was observed consistently that model prediction of DP/DT the work reported in this paper.
deviated from measured DP/DT. It was obvious that measurements of
pipe wall temperature were NOT at all representative of the average Acknowledgements
temperature along the length of the respective pipe segments. In
these two tests, only one pipe wall temperature was measured at one This paper is part of a research program sponsored by TC Energy,
extreme end of the respective piping assembly that clearly did not Calgary, Alberta, Canada, and permission to publish is gratefully
provide accurate indication representing the entire length of the acknowledged.
system.
7. The present work used ANSYS FEA methodology to solve the perti­ References
nent thermal transient governing equation of both buried and above
ground pipe segments. In this model, it is assumed that the convec­ AN INTERNATIONAL EXECUTIVE POWER SECRETARY of ENVIRONMENT and
NATURAL RESOURCES Official Mexican STANDARD NOM-007-ASEA-2016,
tive motion of the test medium inside the pipe due to temperature Transportation of Natural Gas, Ethane and Gas Associated with Coal Ore through
gradient (i.e. buoyancy driven flow) is neglected. While this is a Pipelines.
reasonable approximation, future work should account for possible ANSYS Mechanical, 2020. Release.
ASME B31.3, 2018. (Revision of ASME B31.3-2016) Process Piping. ASME Code for
convective fluid motion via CFD analysis to determine the relative Pressure Piping, p. B31.
accuracy of FEA vs. CFD results, particularly when a previous study ASME B31.4, 2019. (Revision of ASME B31.4-2016) Pipeline Transportation Systems for
by Lanzafame et al. (2017) (Lanzafame et al., 2017) indicated that Liquids and Slurries. ASME Code for Pressure Piping, p. B31.
ASME B31.8, 2018. (Revision of ASME B31.8-2016) Gas Transmission and Distribution
CFD analysis outperforms FEM in capturing the internal convective
Piping Systems. ASME Code for Pressure Piping, p. B31.
fluid motions. Future work should also address the optimum number Bahadori, A., Vuthaluru, H.B., 2009. Prediction of bulk modulus and volumetric
of temperature measurements to be taken during hydrotesting along expansion coefficient of water for leak tightness test of pipelines. Int. J. Pres. Ves.
Pip. 86, 550–554. https://doi.org/10.1016/j.ijpvp.2009.01.007.
the pipe section, and the number and strategic location(s) depending
CSA Z662:19, National Standard of Canada, Oil and Gas Pipeline Systems.
on the piping layout, particularly length, geometry, and surrounding DOE - 49 CFR Ch. I (10–1–11 Edition), 192—TRANSPORTATION of NATURAL and
environment. OTHER GAS by PIPELINE: MINIMUM FEDERAL SAFETY STANDARDS.
DOE- 49 CFR Ch. I (10–1–11 Edition), PART 195—TRANSPORTATION OF .HAZARDOUS
LIQUIDS BY PIPELINE.
Credit authors statement Guignon, B., Aparicio, C., Sanz, P.D., 2010. Volumetric properties of pressure-
transmitting fluids up to 350 MPa: water, ethanol, ethylene glycol, propylene glycol,
K.K. Botros: Principal Investigator, Problem Formulation, Governing Castor oil, silicon oil, and some of their binary mixture. J. Chem. Eng. Data 55 (9),
3017–3023.
Equations, Model Development, Data curation, Formal analysis, Inves­ Hartono, A., Kim, I., 2004. Calculation of vapor-liquid equilibria for methanol-water
tigation, Methodology, Software, Validation, Visualization, Writing - mixture using cubic-plus-association equation of state, trondheim. http://www.nt.
original draft, Writing - review & editing, Final Manuscript. J. Crowe: ntnu.no/users/haugwarb/KP8108_Phase_Equilibria/Essays/ardi_hartono_and_inn
a_kim.pdf.
Data curation, Formal analysis, ANSYS Simulations, Validation, Visu­ Holman, J.P., 2009. Heat Transfer, tenth ed. McGraw-Hill.
alization, Writing - review & editing, Writing - review & editing. V. Liu: https://shop.standards.govt.nz/catalog/2885.5:2012(AS%7CNZS)/scope.
Conceptualization, Technical Guidance, Field Experience, Data https://www.saiglobal.com/PDFTemp/Previews/OSH/AS/AS3000/3700/3788-2006.
pdf.

12
K.K. Botros et al. Journal of Natural Gas Science and Engineering 92 (2021) 103989

Huc, M., Zakelj, G., Urbic, T., 2015. Properties of methanol-water mixtures in a coarse- Lemmon, E.W., Huber, M.L., McLinden, M.O., 2010. NIST Standard Reference Database
grained model. Acta Chim. Slov. 3, 524–530. 23: Reference Fluid Thermodynamic and Transport Properties - REFPROP. National
Khatami, A., Kundral, S., VanGennip, K., 2020. Pass/Fail Criterion for HDPE Pipe Institute of Standards and Technology, Standard Reference Data Program,
Pressure Testing Using Incompressible Fluid. SCS Eng. (n.d.). https://www.scseng Gaithersburg, Version 9.0.
ineers.com/scs-advice-from-the-field-how-to-compensate-for-the-effect-of-the-amb Matta, L., 2017. Collective Effects of Leakage, Temperature Changes, and Entrapped Air
ient-temperature-variations-on-the-pressure-changes-within-the-pipe-during-hdpe- during Hydrostatic Testing. Pipeline Pigging and Integrity Management Conference,
pipe-pressure-testing-using-incompressible-fluid/. (Accessed 25 November 2020). Houston, TX, USA.
Kirkwood, M.G., Cosham, A., 2000. Can the pre-service hydrotest be eliminated. Pipes A Procedure for the Hydrostatic Pressure Testing of Marine Facility Piping, 2003. In:
Pipelines Int. 45, 5–19. https://www.coursehero.com/file/60697432/California-hydro-testing-procedure
Kubota, B.Y.H., Tsuda, S., 1979. Specific volume and viscosity of methanol-water pdf/.
mixtures under high pressure. Rev. Phys. Chem. Jpn. 49, 59–69. Sagdeev, D.I., Fomina, M.G., Abdulagatov, I.M., 2017. Density and viscosity of propylene
Kume, D., Sakoda, N., Uematsu, M., 2005. An Equation of State for Thermodynamic glycol at high temperatures and high pressures. Fluid Phase Equil. 450 (25), 99–111.
Properties for Methanol, Seventeenth Eur. Conf. Thermophys. Prop. , 5 - 8 Sept. Soetens, J., Bopp, P., 2015. Water–methanol mixtures: simulations of mixing properties
2005. Bratislava, Slovak Repub. http://www.ectp.sav.sk/. over the entire range of mole fractions. He J. Phys. Chem. B. 119, 8593–8599.
Kunz, O., Wagner, W., 2012. The GERG-2008 wide-range equation of state for natural Zhuravlev, V.I., 1992. Structure of multiatomic alcohols and their solutions according to
gases and other mixtures: an expansion of GERG-2004. J. Chem. Eng. Data 57 (11), the dielectric-spectroscopy data - equilibrium and dynamic properties of
3032–3091. propanediols. Zh. Fiz. Khim. 225–236.
Lanzafame, R., Mauro, S., Messina, M., Brusca, S., 2017. Heat exchange numerical
modeling of a submarine pipeline for crude oil transport. Energy Procedia 126,
18–25. https://doi.org/10.1016/j.egypro.2017.08.048.

13

You might also like