Вы находитесь на странице: 1из 18

Applied Mathematical Modelling 40 (2016) 10020–10037

Contents lists available at ScienceDirect

Applied Mathematical Modelling


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

Steady-state non-isothermal flow model for natural gas


transmission in pipesR
Alfredo López-Benito a,∗, Francisco J. Elorza Tenreiro a, Luis C. Gutiérrez-Pérez b
a
Universidad Politécnica de Madrid, Department of Geological and Mining Engineering, Ríos Rosas, 21 28003 Madrid, Spain
b
ENAGAS, S.A., Department of Technology and Innovation, Autovía A-2, km. 306,4 50012 Zaragoza, Spain

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

Article history: In this article, we describe the work carried out for developing an accurate and fast nu-
Received 24 June 2014 merical method that can efficiently integrate the mathematical model of natural gas flow
Revised 20 June 2016
in a pipeline under non-isothermal steady-state conditions. To do this, we transform the
Accepted 29 June 2016
model equations in order to obtain a system of three explicit expressions for the spatial
Available online 9 July 2016
derivatives of gas velocity, pressure, and temperature, which can be easily integrated us-
Keywords: ing any suitable numerical method. The proposed method enables us to determine the gas
Natural gas transport velocity, pressure, and temperature anywhere in the pipeline. Furthermore, we develop
Gas dynamics two procedures based on this method, one to calculate the flow rate and the other to ex-
Computational fluid dynamics perimentally determine the friction factor. We illustrate the performance of the proposed
Numerical simulation method using several numerical exercises and we compare the results with field measure-
ments on a Spanish pipeline (provided by the company Enagas) and with those obtained
using other methods.
© 2016 Elsevier Inc. All rights reserved.

1. Introduction

Natural gas distribution networks should be able to transport gas from the supply points to the delivery points even if
there are occasional breakdowns in the flow of gas in some of their branches. This means that the network must have a
number of redundant pipelines, and as a consequence, its topological complexity is usually high with pipelines having very
different degrees of filling.
Generally, the simulation of these types of networks is a computationally intensive process [1], especially when a high
degree of accuracy is required, because of the usually high dimensionality of the simulated networks. Therefore, the numer-
ical methods used should be very stable in any operating situation of the network, and it should allow the use of coarse
computational grids without loss of accuracy to minimize the calculation time.
A well-known mathematical model [1] for natural gas flow through a pipeline results from applying the principles of
conservation of mass, momentum, and energy to a cylindrical pipe section of infinitesimal length. It is a system of im-
plicit non-linear differential and algebraic equations relating gas velocity, pressure, temperature, and internal energy with a

R
This article is a result of the collaboration between Enagas, S.A. and the Universidad Politécnica de Madrid for the development of the PipeCalc code
for the numerical simulation of natural gas flow in pipelines.

Corresponding author.
E-mail addresses: alfredo.lopez@upm.es (A. López-Benito), franciscojavier.elorza@upm.es (F.J. Elorza Tenreiro), lcgutierrez@enagas.es (L.C. Gutiérrez-
Pérez).

http://dx.doi.org/10.1016/j.apm.2016.06.057
0307-904X/© 2016 Elsevier Inc. All rights reserved.
A. López-Benito et al. / Applied Mathematical Modelling 40 (2016) 10020–10037 10021

number of thermodynamic and design parameters. In addition, the design of a gas distribution network involves the place-
ment of a number of non-pipe elements, such as compression and regulation stations and control valves, which must also
be modeled. Depending on the purpose of the study, the variables of interest may be the gas pressures and temperatures at
different points of the pipeline or the flow rate (mass or volume) of the flowing gas. If the pipeline is linear, i.e., it has no
loops or points belonging to more than one branch, the most common problem is to determine the gas temperatures and
pressures at a number of locations along the pipeline for a known gas flow rate. In contrast, if the network is more complex,
with loops and branches, or if the capacity of a pipeline is required to ensure that the pressure drop does not exceed a fixed
value, it is necessary to determine the flow rate as a function of pressures and temperatures at the network nodes.
Unfortunately, these mathematical models lack an analytical solution. A first approach is to introduce a number of sim-
plifying assumptions in the model, e.g., isothermal regime and negligible inertial terms. This procedure leads to explicit
expressions relating pressure and flow, such as the Weymouth or Panhandle formulae (see, e.g., [1–3]). These formulae and
other similar ones, far from being obsolete, are today implemented in many commercial simulators, and are widely used in
recent works of pipeline design (see, e.g., [4–9]).
Nowadays, the use of numerical methods for solving the gas flow model has been adopted as a line of research by
numerous research groups. Their aim is to obtain realistic and accurate predictions for the response of a pipeline to different
operating conditions. Some of them avoid the use of the energy-balance equation, considering only the isothermal flow case,
and they use numerical methods adapted to the strong hyperbolic character of equations. Thus, hybrid TVD schemes are
used in [10], while in [11–14], the use of second-order relaxed schemes [15] is preferred, and in [16], the proposed method
is a time-space least-squares spectral method.
A comprehensive approach to the modeling and simulation of gas transmission is described both quantitatively and qual-
itatively in [17]. In that paper, the time-dependent fluid dynamic equations are scaled to obtain a dimensionless model. This
procedure enabled the authors to develop simplified models adapted for each fluid situation using appropriate asymptotic
expansions. The well-posedness of the problem, particularly at the junctions of pipes, has been the subject of several papers
such as [11,18].
In [2] it was reported that there exists a significant difference in the pressure profile along the pipeline between isother-
mal and non-isothermal processes. In that article, and in [19,20], the numerical solution is carried out using the method
of lines [21] together with a five-point biased upwind scheme, after making some simplifications in the energy equation. A
review of commercially available simulation software can be found in [22].
Although pipeline operations are inherently transient processes, steady-state models are widely used to design pipelines
and to estimate the flow and the line pack [23]. Steady-state models are typically the core models for optimizers, schedulers,
and marketing tools [24]. Moreover, a very popular model used by the engineering community for modeling the transient
regime is the so-called quasi-static model, which is essentially a succession of steady-states [14].
The main objective of the present paper is to describe a procedure to carry out quantitative analysis of the gas flow
using the full gas flow model. The first methodological goal is to provide a new expression for the gas flow model in
which the velocity, pressure, and temperature derivatives appear as explicit functions of the velocity, pressure, temperature,
and flow-rate values. Once this is done, the integration of the model equations can be performed using the classic Runge–
Kutta method. Depending on the imposed boundary conditions, this method enables the prediction of the gas pressures,
temperatures, or flow-rate at every point along a pipe. The model is valid only for simulating the steady state in fully
buried terrestrial buried pipelines. Furthermore, it can be used to experimentally determine the friction factor when the
pressures at both ends of the pipeline and the gas flow rate are known.
This article focuses on the numerical model of gas flow through a pipe, leaving the inclusion of non-pipe elements as
the subject of future studies.

2. Mathematical model

The model used in the present work describes the behavior of the gas flowing through a fully buried pipe at a depth
H in the ground, as a function of a spatial coordinate x measured along the pipe axis. Only piecewise straight pipes with
a constant circular cross-section of internal diameter D are considered. The elevation of the pipe, z, is measured vertically
from an arbitrary horizontal plane to the pipe axis. Therefore, the angle, θ , formed by the pipe axis and the horizontal plane
is such that sin θ = z/x (see Fig. 1).
In the context of this study, a control volume (CV) is a small segment of pipe having length x (see Fig. 1 again). The
forces acting on the CV are those due to the shear stress, τ , and to the gas pressure, p, and the gravity force, G. Because the
length of a pipe is always much greater than its diameter, the gas flow can be adequately described using a one-dimensional
(1D) approach. Furthermore, it will be considered that the pipeline diameter does not change with time, i.e., no pipeline
dilatation effects will be taken into account. The application of the principles of conservation of mass, momentum, and
energy to the gas contained in the CV gives the following system of differential equations for the steady-state regime (see
[1,25,26] among others):

d (ρ u )
=0 ⇒ ρ u = constant, (1)
dx
10022 A. López-Benito et al. / Applied Mathematical Modelling 40 (2016) 10020–10037

Fig. 1. Control volume (CV) in a gas pipe.

dp du ρ u2 λ dz
= −ρ u − − gρ , (2)
dx dx 2 D dx

d
 p
 4qw dz
ε+ =− −g , (3)
dx ρ Dρ u dx

where ρ is the gas density, ε is the specific energy (internal energy plus kinetic energy per unit mass), u is the gas velocity,
λ is the Darcy’s friction factor (8τ /ρ u2 ), and qw is the external wall heat flux. Because gas flow is turbulent, with a Reynolds
number typically greater than 105 , we considered that all variables and thermodynamic properties appearing in (1)–(3)
are fairly uniform over the pipe cross-section, except very close to the wall (boundary layer), so u, p, T, ε , and ρ in these
equations represent the averaged values over the cross-section of the CV for the limit x → 0. Only qw is averaged along
the perimeter of the CV cross-section, and it will generally depend on x.
Eq. (3) is a usual form of the equation of conservation of energy. However, it may take other forms depending on how the
thermodynamic properties are managed. For instance, the use of the specific bulk enthalpy, h = ε + p/ρ − u2 /2, in (3) yields:

d
1 4qw

dz
h + u2 = − −g (4)
dx 2 Dρ u dx

Furthermore, the equation of state is the relationship between the pressure, temperature, and density:

p = ρ RT Z ( p, T ), (5)

where R is the specific gas constant, T is the temperature, and Z is the compressibility factor.
Finally, in order to obtain a unique solution of the system of Eqs. (1), (2), (4), and (5), initial and boundary conditions
must be specified.

2.1. Heat transfer

In most cases, conduction is the only important mechanism by which heat leaves the pipeline through the pipe walls
(see [27,28]). Only in the case of pipelines that are in contact with water (underwater pipelines) do other effects such as
free convection become important. In this study, we consider only fully buried terrestrial pipelines. The heat conduction in
any material can be described by Fourier’s law, which states:

q = −κ∇ T , (6)

where q is the heat flux, κ is the thermal conductivity of the material, and ∇ T is the temperature gradient. The heat transfer
rate, Q˙ , from an isothermal surface (area A1 ) at temperature T1 to an isothermal surface (area A2 ) at temperature T2 can be
obtained integrating the Eq. (6) over the surfaces. If κ is constant the heat transfer rate can be written as:

Q˙ = κ S (T1 − T2 ), (7)
A. López-Benito et al. / Applied Mathematical Modelling 40 (2016) 10020–10037 10023

where S is the shape factor defined by:


 
A1 (∂ T /∂ n )1 dA1 A2 (∂ T /∂ n )2 dA2
S= =− . (8)
T2 − T1 T2 − T1
Here ∂ T/∂ n are the local temperature gradients over the surfaces and ∂ /∂ n denotes derivatives in the normal direction
pointing into the space between the surfaces [29]. In most cases, integrals appearing in Eq. (7) can only be solved analytically
introducing a number of simplifications.
In steady state and for thermally uninsulated steel pipes, a fair assumption is that the temperature throughout the entire
pipe wall is equal to the gas temperature. This is realistic because the thermal conductivity of steel is much higher than
that of the surrounding ground soil (about 50 W m−1 K−1 for steel, less than 3.5 W m−1 K−1 for ground), the wall thickness
is only a few millimeters and the Reynolds number is high.
Usually, the pipe inclination is small (less than 3◦ ) and the burial depth is almost constant, so we have supposed that
the control volume is a horizontal cylinder of length x buried at a depth H(x), taken to be constant over the control
volume located at position x. Furthermore, the heat transfer inside the pipe is due to convection which is a more efficient
mechanism than conduction, so conduction in the axial direction is neglected and, consequently, heat transfer between gas
and ground soil is studied separately for each cross-section of the pipe.
From these simplifications, the shape factor for the heat transfer from the outer lateral surface of the pipe (at temperature
T) to the ground surface (at temperature Tgr ) calculated for a pipe control volume of length x [29,30] is:
2 π x
S=   , (9)
2H

2H 2
log D+2t
+ D+2t
−1

where H is the pipe-axis burial depth at location x along the pipe and t is the pipe wall thickness. Denoting the thermal
conductivity of the ground soil by κ gr , Eq. (7) for the control volume yields the mean heat flux at pipe wall:

Q˙ κgr S(T − Tgr ) 2κgr (T − Tgr )


qw = = =   , (10)
π (D + 2t )x π (D + 2t )x 
2H 2
(D + 2t ) log 2H
D+2t
+ D+2t
−1

For insulated pipes, Eq. (10) is not valid because the temperature variation in the insulating layer must be accounted for.

2.2. Friction factor

The Darcy–Weisbach friction factor, λ, is a dimensionless number defined as four times the ratio of the local shear stress
to the local flow kinetic energy density:
τ
λ=4 , (11)
ρ u2
2

The friction factor depends on the Reynolds number (Re ) and on the pipeline relative wall roughness (k/D), being stronger
than its dependency on k/D as the Reynolds number increases [22]. In the cases considered in this paper, the Reynolds
number values are near 107 (fully turbulent regime), so it is important to obtain an accurate value of k/D. The problem is
that the surface roughness does not have a generally accepted definition, and it is not easily measured. The only way to
obtain an accurate value for this parameter may be to experimentally measure the pressure loss, and then calculate the
roughness that corresponds to it. The proposed method can be used to obtain the value of this experimental friction factor
in the way described in Section 3.3. However, when experimental measures are not available, a realistic friction factor value
may be estimated using some of the many correlations that have been proposed for this purpose, such as Nikuradse in 1933
[31], Colebrook in 1939 [32], Uhl et al. in 1965 [33], Chen in 1979 [34], Haaland in 1983 [35], GERG 1.19 in 20 0 0 [36], and
Bratland in 2009 [22]. All of them are empirical, and they have their own limits of application depending on the Reynolds
number. In this study, we used the GERG 1.19 correlation because it is the one usually employed by Enagas, following the
GERG (Groupe Europeen de Recherches Gazières) recommendation. It is given as:

 n  0.942n f
1 2 k/D 1.499
√ = − log10 + √ (12)
λ n 3.71 f Re λ

Similarly to other correlations as in the study by Colebrook [32], in (12), the first term in brackets relates to rough pipe
flow (Re → ∞), while the second term relates to smooth pipe flow (k = 0). The novel aspect of this formula is the addition
of the two parameters n and f to consider not only the energy losses due to the wall roughness, which depends only on the
height of the pipe-wall imperfections, but also those due to other pipe features.
It is well known that pipe joints, welds, and fittings, such as valves, cause energy losses that do not depend on the
Reynolds number, and which can be interpreted as an increment of the relative roughness value. Hence, k/D can be used for
tuning the friction factor in the fully rough flow regime (Re → ∞).
10024 A. López-Benito et al. / Applied Mathematical Modelling 40 (2016) 10020–10037

Fig. 2. Effect of parameters n and f on a k/D = 2 × 10−4 pipe.

Furthermore, the pipe geometry causes a type of energy loss that depends on the Reynolds number. Parameter f, which
is called the drought factor, is used to consider this additional loss in the smooth flow regime (k = 0), and its value is always
less or equal to 1.
Fig. 2 shows the effect of these parameters on the Moody-like diagram corresponding to the correlation (12) for a k/D =
2 · 10−4 pipe. Parameter n controls the size of the transition zone between the fully rough and smooth flow regimes. For
n = 1, the transition is very smooth (very much like the situation described by the original Moody diagram [32]), and it
becomes more abrupt as the n value increases. For n = 10, the transition zone is very small, and is very similar to the one
used by AGA [33]. It appears that the transition zone size is related to the geometric distribution of the wall imperfections
[36]. In the latter article, the value n = 10 was recommended.
As (12) is a non-linear implicit equation, the calculation for λ implies the use of a method such as a fixed-point or some
variation of the Newton method (see [22] for details). Either of these two methods converges very quickly (usually in less
than ten iterations). The gas dynamic viscosity, η, which is required to calculate the Reynolds number at each gas pressure
and temperature, can be estimated using some theoretical methods (see [37]). In this paper, we used the method proposed
by Lucas [37].

3. Integrating the model equations

As mentioned in the Introduction section, the gas flow model equations do not have an analytical solution, and we
therefore need to develop a suitable numerical procedure to simultaneously solve Eqs. (1)–(3), and (5) in their original
form, avoiding further model simplifications. This is the goal of the present section.

3.1. Isothermal regime

This is the initial assumption of the most popular formulae relating pressure and flow rate that can be found in classical
text books such as Weymouth and Pandhandle. For the isothermal flow Eq., (3) is redundant, and therefore, the gas flow
model reduces to Eqs. (1) and (2). Furthermore, from Eq. (1):
du u dρ
=− , (13)
dx ρ dx
and so
du u 1 dp
=− , (14)
dx ρ RT Z dx

from Eq. (5) for constant RTZ. Note that RT Z is the velocity of sound in the gas stream, which in practice and at usual
operating conditions, has very little variation along the pipe.
Inserting (14) into (2), the momentum equation can be written in the form

dp RT Z ρ u2 λ dz
=− − gρ . (15)
dx RT Z − u2 2 D dx
A. López-Benito et al. / Applied Mathematical Modelling 40 (2016) 10020–10037 10025

In most countries, gas velocity is limited by law to 20 m/s to prevent excessively high noise during its transport, and in
practice, it seldom exceeds 5 m/s. Of course, the speed of sound is much higher (usually exceeding 300 m/s) and therefore,
in most cases, the assumption of RT Z/(RT Z − u2 ) = 1 is a good approximation. This is equivalent to neglecting the inertia
term ρ u du/dx in Eq. (2).

3.2. Non-isothermal regime

From the perspective of a pipeline designer, the system temperature is a much more useful variable than the enthalpy, so
an expression for dT/dx should be found. This can be done by expressing the energy Eq. (4) in terms of temperature rather
than enthalpy. Differentiating h with respect to x gives
  
dh ∂h dT ∂h ∂T dp dT dp
= − = cp − c pμ , (16)
dx ∂T p
dx ∂T p
∂p h
dx dx dx

and inserting this result in (4) yields


dT dp du dz 4qw
cp − μc p +u +g =− , (17)
dx dx dx dx Dρ u
where cp is the specific isobaric heat capacity and μ is the Joule–Thomson coefficient.
Because point estimates of functions u(x), p(x), T(x), and ρ (x) should be obtained from the system formed by the differ-
ential and algebraic Eqs. (1), (2), (5), and (17), which relate u, p, T, dp/dx, and dT/dx, the first task is to find an expression for
du/dx. This can be done by multiplying Eq. (5) by u, and taking logarithms and differentiating it with respect to the spatial
coordinate, yielding
du dT dp
= Au + Bu , (18)
dx dx dx
with
1 1 ∂Z 1 ∂Z 1
A= + and B= − .
T Z ∂T Z ∂p p
Once this last equation has been obtained, the system (1), (2), and (17) can be rewritten as
du u[A(c1 + c2 μc p ) + c2 Bc p ]
= , (19)
dx C

dp Au2 (ρ c1 − c2 ) − c2 c p
=− , (20)
dx C

dT Bu2 (ρ c1 − c2 ) + c1 + c2 μc p
= , (21)
dx C
with
 
C = Au2 β T + c p 1 + ρ Bu2 , (22)

4qw dz
c1 = − −g ,
Dρ u dx
ρ u2 λ dz
c2 = − − gρ ,
2D dx
where β is the thermal expansion coefficient, which is defined as

∂ρ 1 1 + ρ c pμ
β=− = .
ρ ∂T p
T

Note that Eq. (19) is not required because once expressions (20) and (21), which do not depend on du/dx, have been
obtained, the gas velocity can be computed from the continuity Eq. (1). The data input is often not the mass flow rate, but
it is instead given in terms of the volumetric flow rate in given conditions Tb and pb . Usually, these given conditions are
the standard conditions (STP), where Tb = 273.15 K and pb = 1.01325 bar, or the normal conditions (NTP), where Tb = 293.15 K
and pb = 1.01325 bar. In either case, from Eq. (1) for a constant cross-sectional area A = π D2 /4.
4 Q b ρb
u= , (23)
π D2 ρ
where Qb and ρ b are respectively the volumetric flow rate and the gas density at conditions T = Tb and p = pb . Eq. (23) can
replace Eq. (19).
10026 A. López-Benito et al. / Applied Mathematical Modelling 40 (2016) 10020–10037

In addition, three calculations have to be performed at each computational step. These are the friction factor, which is
calculated using (12), the natural gas properties, which are calculated from the gas composition at a given pressure and tem-
perature using the ISO-WD 20765–1 norm [38], and the gas dynamic viscosity (used for the Reynolds number calculation),
which is determined using the method proposed by Lucas [37].
Eqs. (23), (20), and (21) relate the gas pressures and temperatures at the pipe ends (p1 , p2 , T1 and T2 ), the gas flow rate
at standard conditions (Qb ), and the gas velocity (u), of which three must be specified to make the system solvable. Here,
we examine specific cases.

3.2.1. First case: calculate the gas pressure, velocity, and temperature along the pipeline, knowing the flow rate and the inlet
pressure and temperature
If the conditions of gas pressure and temperature are known at the inlet of the pipe, the density at this point can be
estimated from the equation of state, and the problem reduces to solving an initial-value problem, where Eq. (23) provides
the gas velocity along the pipe, and it is only required to integrate Eqs. (20) and (21) along the pipe. Any conventional
numerical method for solving initial-value problems can be used for this purpose. Here, a classical fourth-order Runge–
Kutta method (see [39,40]) is preferred because it has a good balance between accuracy and computational cost.

3.2.2. Second case: calculate the volume flow rate from the pressures and temperatures at the ends of the pipe
As the pressure, p2 , and temperature, T2 , at the outlet are obtained from the pressure, p1 , and the temperature, T1 ,
respectively, at the inlet and the flow-rate under standard conditions, Qb (see 3.2.1), we can assume the existence of two
functions such as
p2 = Fp (Qb , p1 , T1 ) (24)

T2 = Ft (Qb , p1 , T1 ) (25)
Eqs. (23)–(25) form a system of three equations linking six variables: u, Qb , p1 , p2 , T1 , and T2 , so it is necessary to fix
three of them.
If p1 , T1 and Qb are the fixed variables, the resolution of this system can be carried out using the method explained in
Section 3.2.1, but if other variables are fixed, a method suitable for non-linear systems of equations has to be used. From
the numerical experiments carried out on these equations, it has been deduced that the functions Fp and Ft are smooth,
continuous, and have only one root at usual operating conditions. Thus, the system (24) and (25) can be solved by the
Newton’s method using a numerical approximation for the derivatives of the Jacobi matrix.
A commonly occurring scenario exists when simulating transmission networks, where considering the network as a
graph, the system of equations AQ = L must be solved. In this system, A is the nodal incidence matrix of the graph, Q
is a vector containing the flow rates at the network pipes (branches of the graph), and L is a vector representing the nodal
loads (deliveries or supplies). This system represents Kirchhoff’s first law, which states that the sum of the mass flow rates
at a node (or equivalently, the sum of the volume flow rates under standard conditions) is zero. In addition, we considered
that at each node, there is a unique pressure, and that the following relationship is fulfilled:

j∈δi− Q j T j−
= Tk+ , ∀k ∈ δi+ , (26)
j∈δi− Qj

where δi− (δi+ ) denotes the set of all indices j of pipes going in (out) of the node i, Qj is the volume flow rate at pipe j (at
standard conditions), and T j− (T j+ ) is the temperature at the outgoing (ingoing) node of the pipe j.
The system constituted by AQ = L and Eqs. (26), plus the condition of uniqueness of pressure at nodes, is non-linear,
and it is usually solved by Newton’s method for the nodal values of pressures and temperatures. For each iteration, a set
of nodal values of pressure and temperature is obtained, and from these values, the flow rates at each pipe are calculated.
Hence, at each iteration and for each pipe, a system such as (24) and (25) is solved.
Although Newton’s method is relatively robust and fast, to ensure a good convergence, it is desirable to use a starting
value that is not too far from the solution. A good choice is to use the solution obtained by some classical formula. In this
work, we used the Pandhandle B equation [1,3]. In addition, the gas velocity asymptotically increases with Qb , making it
necessary to put controls in the algorithm in order to keep the gas velocity under reasonably limits, say 30 m/s, during the
iteration.

3.2.3. Experimental determination of the friction factor from measurements of gas pressure, temperature, and volume flow rate
As stated earlier, the uncertainty in determining the friction factor is one of the most important reasons for inaccurate
results, and it can only be reliably determined from experimental data. Here, we present a method that can be used to
perform this task. In Section 3.2.2, it was assumed that the friction factor is known. However, if λ is considered as an
unknown, and p1 , p2 , T1 , T2 , and Qb are known, Eqs. (27) and (28) become:
Fp (λ ) − p2 = 0, (27)

Ft (λ ) − T2 = 0, (28)
A. López-Benito et al. / Applied Mathematical Modelling 40 (2016) 10020–10037 10027

which, assuming constant λ along the pipe, can be used to calculate the friction factor in an entirely analogous way to that
used to determine Qb .
Once the friction factor has been determined, it is possible to calibrate the parameters that appear in the empirical
correlations as the roughness or the drought factor.

4. Simplified models

At normal gas pipeline operation conditions, the contribution of the inertia term (ρ u du/dx) in Eqs. (2) and (17) is very
small (see [1,22] among others), and it is very often neglected. Moreover, the compressibility factor depends mainly on
temperature because the term ∂ Z/∂ p is of the order of 10−7 , so in most practical circumstances, we can assume B = 0. If
these simplifications are made, model (19)–(21) reduces to
ρ u = constant, (29)

dp
= c2 , (30)
dx
dT c 1 + μc p c 2
= . (31)
dx Au2 + c p
Based on this simplified model, it is possible to obtain three models similar to those obtained by Brower et al. [17], who
considered the following special cases:

1. If the heat loss through the pipe wall is negligible, then qw can be regarded as zero, and hence, the value of the coeffi-
cient c1 can be taken as c1 = −g dz/dx.
2. If the heat loss through the pipe wall is important, then the full model (29)–(31) must be considered.
3. At points of the pipe away from sources of hot gas such as the outlets of compressors, the gas is nearly in thermal
equilibrium with the ground surrounding the pipe, and in practice, an isothermal regime can be assumed. In this case,
the energy equation is removed from the model. Note that when this occurs, the energy losses are only due to the
internal pipe-wall friction and gravity.

These reduced formulations may be suitable for many practical applications.

5. Applications

The numerical method proposed in this article was tested by solving numerous cases [41], some of which will be shown
below. In the first case studied, the values of gas pressure and temperature as well as the friction factor predicted by
the proposed method were contrasted with field measurements performed on a pipeline located in western Spain by the
company Enagas. Values of the measured gas pressures, temperatures and deliveries have been included in the sketch map
of this pipeline displayed in Fig. 3. In addition, Table 1 shows the gas composition, which is required by the ISO-20765-1,
and the Lucas method for calculating the gas properties.
We also investigated the performance method by comparing it with other similar methods described in the literature.
The methods chosen for comparison are those described by Osiadacz & Chaczykowski [2] (hereafter referred to as the AGA-
Osiadacz method) and Abdolahi et al. [42] (hereafter referred to as the Abdolahi method). We chose the AGA-Osiadacz

Fig. 3. Case study 1: sketch map of the pipeline with the measured data, where z is the node elevation and θ is the angle formed by the pipe axis and
the horizontal plane (see Fig. 1). Flow-rates are given at standard conditions.
10028 A. López-Benito et al. / Applied Mathematical Modelling 40 (2016) 10020–10037

Table 1
Gas composition used in the applications.

Component Mole fractions

Methane C1 0.8686
Nitrogen N2 0.0106
Carbon dioxide CO2 0.0181
Ethane C2 0.0844
Propane C3 0.0153
i-Butane i-C4 0.0012
n-Butane n-C4 0.0014
i-Pentane i-C5 0.0 0 02
n-Pentane n-C5 0.0 0 01
n-Hexane n-C6 0.0 0 01

Table 2
Results for case study 1. These data are plotted in Fig. 4.

Node Measured Predicted Difference

p (bar ) T ( C)

p (bar ) T ( C)◦
p (bar ) T (◦ C )

1 78.3 26.00 78.300 26.0 0 0 0.0 0 0 0.0 0 0


2 76.2 25.10 75.944 24.812 0.256 0.331
3 75.4 25.00 75.029 24.860 0.371 0.182
4 74.5 24.80 74.124 24.699 0.376 0.142
5 74.0 24.80 73.667 24.730 0.333 0.108
6 73.3 24.60 73.048 24.683 0.252 −0.048
7 71.6 23.60 71.397 23.664 0.203 −0.022
8 68.6 22.00 68.337 21.656 0.263 0.404
9 67.0 21.70 66.986 21.232 0.014 0.549

The friction factor was calculated using the GERG 1.19 formula.

method primarily because it is a simple and widely used non-differential procedure, and we chose the Abdolahi method
because it is similar to that proposed here, and it allows us to show how the extra effort to reduce to two the number
of differential equations of the model and to make it explicit is rewarded by a significant enlargement of the grid-size for
which it is numerically stable. The key difference between the method proposed by Abdolahi et al. and the proposed method
is the procedure for performing the numerical integration of the model. At each point of the computational grid, they solved
the system (2), (17), and (18) using Cramer’s method at each step of the classical fourth-order Runge–Kutta method in order
to obtain the values of du/dx, dp/dx, and dT/dx. The method proposed here is fully explicit, and requires the integration of
only two differential equations because the velocities are calculated from the equation of conservation of mass, which is
algebraic in the steady state.
Before describing these case studies, we need to clarify the concept of node. In the following paragraphs, the pipeline will
be described as a set of points, named nodes, linked by pipes. In other words, a “node” is a point at which some property
of the pipeline, such as the inner diameter and slope, abruptly changes, or simply a point at which measurements for the
gas pressure, temperature, and flow-rate are available.

5.1. Case study 1: comparison of the predicted gas temperature, pressure, and friction factor values with measured data

In the first case to be considered, we computed the pressure values using the proposed differential method, and we
compared it with experimental data obtained from a 203-km-long pipeline located in western Spain, which is composed
of eight serial pipes. The inlet node is the outlet of a compressor, and several deliveries have been placed at nodes along
the pipeline (see Fig. 3. Additional data obtained are the roughness, k = 12 μm (provided by the manufacturer), internal
pipe diameter, D = 0.6426 m, wall thickness, t = 8.9 mm, depth of pipe axis below the ground surface, H = 1.5 m, and the
thermal conductivity of the ground soil, κgr = 3.3 Wm−1 K−1 . We performed the simulation using the GERG 1.19 formula
(12) to estimate the friction factor and a mean mesh size of 10 0 0 m.
The results for this numerical experiment are summarized in Table 2. In Fig. 4, the measured and predicted values for
pressures and temperatures are plotted against the x-coordinate. In the table and figure, we observe good agreement be-
tween the real and computed pressure values. In fact, the differences between them do not exceed 0.376 bar and 0.549 ◦ C
when the GERG 1.19 correlation for the friction factor is used, and they do not appear to be correlated to the x-coordinate,
highlighting the good behavior of the numerical method. This low error value becomes more relevant if we consider that
measurements at the pipeline have an accuracy of only 0.1 bar and 0.1 ◦ C, respectively, and that the measurements of the
ground temperature and thermal conductivity are scarce and inaccurate. Moreover, the data are obtained from field mea-
surements, which are much less accurate than laboratory measurements. Another factor that must be considered when eval-
uating the results is that the experimental data represent an instant of the pipeline operation, which is not in steady state.
A. López-Benito et al. / Applied Mathematical Modelling 40 (2016) 10020–10037 10029

Fig. 4. Case study 1: comparison of measured and predicted data. The friction factor was estimated using the GERG 1.19 formula.

Fig. 5. Case study 1: Reynolds number, velocity, and dynamic viscosity calculation obtained using measured data and experimental friction factor values.
The arrows represent the gas deliveries.

In Fig. 5, the Reynolds number, speed, and dynamic viscosity values are represented as a function of the spatial coordinate,
x.
However, as mentioned in Section 2.2, the most important source of uncertainty may be the friction-factor coefficient
estimations. Fig. 6 shows the friction factor values estimated on each pipe using the various empirical formulae referred
in Section 2.2 and by the proposed method from experimental data. Friction factor values estimated by some of the most
used empirical formulae (Colebrook [32], Chen [34], GERG 1.19 [36], and Bratland [22]) range from 0.0089 to 0.0097. Those
obtained using the GERG 1.19 formula appear to be near the central values of this range. However, the friction factor val-
ues obtained from experimental data significantly differ from these estimations. Only in the central part of the pipeline is
there good agreement between measured and predicted values. The values of pm − p p (measured pressure minus predicted
pressure) are depicted in Fig. 7, and they show differences that can exceed one bar.
10030 A. López-Benito et al. / Applied Mathematical Modelling 40 (2016) 10020–10037

Fig. 6. Comparison of experimental and estimated friction factor values.

It is even more important to obtain an accurate friction factor when the gas flow rate in a pipeline is to be calculated
from the gas pressures at its ends. We calculated the volume flow rate using the method described in Section 3.2.3 from
the measured data provided by Enagas. The results obtained are depicted in Fig. 8, which show the predicted values of the
gas volume flow rate obtained using several correlations for the friction factor. We observe that there are differences that
can reach almost 16 % of the measured values.

5.2. Case study 2: comparison with other methods

In this case study, we verified the performance of the proposed method compared with the AGA-Osiadacz method
[2,43] and the Abdolahi method [42]. First, we applied the three methods to predict the gas pressures and temperatures
at the same pipeline described in case study 1. Secondly, we investigated the CPU-times, stability, and grid-size dependence
A. López-Benito et al. / Applied Mathematical Modelling 40 (2016) 10020–10037 10031

Fig. 7. Comparison of the difference between measured and predicted pressure for several correlation formulae of the friction factor.

of these methods on a 500-km-long single horizontal pipe and on a small network composed of four nodes linked by five
pipes. In all simulations carried out, we estimated the friction factor using the GERG 1.19 formula (12).
As can be seen in Fig. 9, the differences between the predicted and measured pressures are small for the three methods;
however, in the case of the temperatures, the AGA-Osiadacz method provides predictions that are significantly different from
those of the other two tested methods.
For the set of simulations on a 500- km-long single horizontal pipe, we used the following parameters: inner pipe di-
ameter, D = 0.6426 m, inlet gas pressure, p = 80 bar, inlet gas temperature, t = 30 ◦ C, and gas flow rate at standard pressure
and temperature conditions, Qb = 80 m3 s−1 . We estimated the friction-factor coefficient using the GERG 1.19 correlation.
Fig. 10 shows the results of these experiments. In this figure, N represents the number of pipe sections that make up the
computational grid.
The first conclusion to be made is that the AGA-Osiadacz method requires less CPU time than the other two methods (see
Fig. 10(a)). In addition, it is numerically stable for any value of N, although the predicted outlet pressure strongly depends
on the value given to this parameter. With respect to the differential methods that were compared, both are numerically
stable for N values greater than 10 (corresponding to a grid size smaller than 50 km).
Once the value of N for which the method is numerically stable is reached, the predicted outlet pressures (see Fig. 10(b))
and temperatures (see Fig. 10(c)), which were computed using the proposed method, are independent of the value of N,
while these values, especially the predicted outlet pressure, depend on the grid size when they are computed using either
of the other two methods.
The differential methods compared here were also tested using the network depicted in Fig. 11, which comprises four
nodes and five pipes. We imposed deliveries of 200 m3 /s and 100 m3 /s (under standard conditions) at nodes 2 and 4, re-
spectively, while at node 1, the amount of gas required to meet the demand was supplied at 80 bar and 40 ◦ C, respectively.
Pipe Pi-01 has a 1-m internal diameter, while the other pipes have 0.5-m internal diameters. The ground soil is at 25 ◦ C and
has a thermal conductivity of 3.3 W m−1 K−1 . Other parameters are t = 7.9 mm, H = 1.5 m, k = 12 μm, n = 10, and f = 0.96
for all pipes. The simulation results are given in Tables 3 and 4.
In both methods, we used a uniform computational grid for each pipe. The grid size, x, is taken to be as the largest
size that ensures that the results obtained for x and smaller grid sizes do not differ by more than 0.001 ◦ C and 0.001 bar.
Tables 3 and 4 show similar results for both methods. However, the key difference is the computation time; for this
application, the proposed method is almost twice as fast as Abdolahi’s method. This is because the proposed method usually
requires a much coarser computational grid to give accurate results, especially for pipes transporting a small amount of gas
as Pi-02.
10032 A. López-Benito et al. / Applied Mathematical Modelling 40 (2016) 10020–10037

Fig. 8. Comparison of the volume flow rates predicted using several correlation for friction factor and their measured values.

Table 3
Comparison of nodal results for network simulation.

Proposed method Abdolahi

Node p L T p L T
(bar ) ( m3 / s ) (° C) (bar) (m3 /s ) (◦ C )

N-01 80.0 0 0 0 −300.0 0 0 0 40.0 0 0 0 80.0 0 0 0 −300.0 0 0 0 40.0 0 0 0


N-02 77.3618 200.0 0 0 0 25.6362 77.3645 200.0 0 0 0 25.6833
N-03 77.2954 0.0 0 0 0 – 77.3016 0.0 0 0 0 –
N-04 73.7661 100.0 0 0 0 24.6871 73.7799 100.0 0 0 0 24.8529

T is the temperature of gas ingoing or outgoing the network.


A. López-Benito et al. / Applied Mathematical Modelling 40 (2016) 10020–10037 10033

Fig. 9. Case study 2: differences between predicted and measured data. In the predictions, we used the GERG 1.19 formula to estimate the friction factor
values.

Table 4
Comparison of pipe results for network simulation.

Proposed method Abdolahi

Pipe Node 1 Node 2 N x Qb T1 T2 N x Qb T1 T2


(km ) (m3 /s ) (◦ C ) (◦ C ) (km ) (m3 /s ) (◦ C ) (◦ C )

Pi-01 N-01 N-02 5 20.00 205.5366 40.0 0 0 0 25.6362 4 25.00 205.3774 40.0 0 0 0 25.6833
Pi-02 N-02 N-03 20 3.54 5.5366 25.6362 24.9996 47 1.50 5.3774 25.6833 24.9998
Pi-03 N-01 N-03 7 10.10 41.4144 40.0 0 0 0 24.8480 12 5.89 41.5045 40.0 0 0 0 24.9199
Pi-04 N-03 N-04 4 17.68 46.9511 24.8659 24.7400 7 10.10 46.8819 24.9291 24.8805
Pi-05 N-01 N-04 8 12.50 53.0489 40.0 0 0 0 24.6403 12 8.33 53.1181 40.0 0 0 0 24.8285

CPU time 33.969 s 65.844 s


10034 A. López-Benito et al. / Applied Mathematical Modelling 40 (2016) 10020–10037

Fig. 10. Case study 2: behavior of three different numerical methods against the computational grid size. (a) CPU time, (b) pressure at the final node
(outflow node), (c) temperature at the final node (outflow node).

In summary, the proposed method has a similar accuracy to that of the other methods investigated. However, their
results are influenced less by the grid size. This may lead to considerably lower computing times when massive calculations
are performed, as is the case with the simulation of large distribution networks.

6. Conclusions

In this study, we developed a numerical method for the prediction of gas pressures, temperatures, and velocities at
any point along a pipeline. In addition, we solved two inverse problems using algorithms based on our proposed method,
namely, the calculation of the flow rate of the gas stream and the experimental determination of the pipe-friction factor.
A. López-Benito et al. / Applied Mathematical Modelling 40 (2016) 10020–10037 10035

Fig. 11. Case study 2: Network used to compare the behavior of the proposed method and Abdolahi’s method.

Table 5
List of symbols and units.

Symbol Unit Description

−1
cp J · kg · K−1 Isobaric specific heat capacity
−1
cv J · kg · K−1 Isochoric specific heat capacity
D m Pipe diameter
−1
e J · kf Specific energy of gas
f – Drought factor in GERG 1.19 correlation
g m · s−2 Gravitational acceleration
h m2 · s−2 Specific bulk enthalpy
H m Depth of pipe axis below ground surface
k m Pipe wall roughness
L m Pipe length
n – Factor in GERG 1.19 correlation
p N · m−2 Gas pressure
qw W · m−2 Mean heat flux at pipe wall
Q˙ W Heat flow rate
Qb m3 · s−1 Volume flow rate at standard conditions
R m2 · s−2 · K−1 Specific gas constant
Re – Reynolds number
S m Shape factor
t m Pipe wall thickness
T K Gas temperature
u m · s−1 Mean gas velocity
us – Geometric factor in Bratland correlation
U W · m−2 · K−1 Overall heat-transfer coefficient
Z – Compressibility factor
β m · K−1 Thermal expansion coefficient
−1
ε J · kg Specific internal energy of gas
γ – Isenthropic exponent
λ – Friction factor
κ gr W · m−1 · K−1 Thermal conductivity of ground soil
μ K · m2 · N−1 Joule–Thomson coefficient
η kg · s−1 · m−1 Gas dynamic viscosity
ρ kg · m−3 Gas density
ρb kg · m−3 Gas density at standard conditions
10036 A. López-Benito et al. / Applied Mathematical Modelling 40 (2016) 10020–10037

The proposed numerical scheme requires a coarser computational grid size to obtain an accuracy that is similar to that
of other methods with which it has been compared (see case study 2). We performed numerical experiments to simulate
the injection of 100 m3 /s (at standard conditions) of gas at 80 bar into 500- km-long horizontal pipes with a diameter
equal or greater than 0.6 m. The results were almost identical for computational grid-size values ranging from 100 m to
50 km. The good performance of the proposed method is even more marked for pipes carrying small amounts of gas, and
this is a significant advantage when massive calculations must be performed, such as in the analysis of large and complex
transmission networks.
When compared with field data, the prediction accuracy depends more on the uncertainties in the underlying empirical
correlations of the model and in the data introduced in it than on the uncertainties that are due to the numerical method.
In this regard, the role played by the factor friction is critical. As can be seen in the cases studied, the differences between
the estimates of this factor performed using the means of empirical correlations and the experimental measurements can
exceed 30 %. As a result of this uncertainty, the errors in the prediction of values of pressure loss can be greater than
12 %, and when predicting the values of the gas flow rate, can exceed 16 %. Compared with these values, the uncertainty
introduced by the choice of the empirical correlation that was used to estimate the friction factor is much smaller.
Finally, the proposed method has been tested with the help of field data which are accurate enough to control the
pipeline operations, but which are obviously much less accurate than data that may be obtained in controlled laboratory
conditions. This is particularly true for the model of heat exchange between the gas and the surrounding ground for which
the available field data for ground temperature and thermal conductivity are very poor.

Symbols and units

In Table 5, there is a list with the symbols and units used in this article. All magnitudes are expressed in units of the
International System of Units.

Acknowledgments

This article includes the work of the research project “PipeCalc: a natural gas transport network simulator development,”
which was financed by the Enagas, S.A. The authors wish to thank this company for the facilities provided and the ongoing
support. We also thank professor J.A. Sanchidrián for his advice on some heat transfer aspects and the anonymous reviewers
whose valuable criticisms and suggestions have significantly helped to improve the original manuscript.

References

[1] A.J. Osiadacz, Simulation and Analysis of Gas Networks, Gulf Publishing Company, 1989.
[2] A. Osiadacz, M. Chaczykowski, Comparison of isothermal and non-isothermal pipeline gas flow models, Chem. Eng. J. 81 (1–3) (2001) 41–51.
[3] D.W. Schroeder, A tutorial on pipe flow equations, in: Proceedings of the Conference on PSIG, PSIG, 2001.
[4] A.D. Woldeyohannes, M.A.A. Majid, Simulation model for natural gas transmission pipeline network system, Simul. Model. Pract. Theory 19 (1) (2011)
196–212.
[5] G. Yan, Y. Gu, Effect of parameters on performance of lng-fpso offloading system in offshore associated gas fields, Appl. Energy 87 (2010) 3393–3400.
[6] J. Kavicky, M. Jusko, B. Craig, E. Portante, S. Folga, A natural gas modeling framework for conducting infrastructure analysis studies, in: Proceedings of
the Winter Simulation Conference (WSC), 2009, pp. 2891–2901.
[7] V.S. Norstebo, F. Romo, L. Hellemo, Using operations research to optimise operation of the norwegian natural gas system, J. Natural Gas Sci. Eng. 2 (4)
(2010) 153–162.
[8] T. Adeosun, O. Olatunde, J. Aderohunmu, T. Ogunjare, Development of unsteady-state Weymouth equations for gas volumetric flow rate in horizontal
and inclined pipes, J. Natural Gas Sci. Eng. 1 (4–5) (2009) 113–117.
[9] Y. Ruan, Q. Liu, W. Zhou, B. Batty, W. Gao, J. Ren, T. Watanabe, A procedure to design the mainline system in natural gas networks, Appl. Math. Model.
33 (7) (2009) 3040–3051.
[10] J. Zhou, M. Adewumi, Simulation of transients in natural gas pipelines using hybrid tvd schemes, Int. J. Numer. Methods Fluids 32 (4) (20 0 0) 407–437.
[11] M. Banda, M. Herty, A. Klar, Gas flow in pipeline networks, Netw. Heterog. Media 1 (1) (2006) 41–56.
[12] M. Banda, M. Herty, J. Ngnotchouye, Toward a mathematical analysis for drift-flux multiphase flow models in networks, SIAM J. Sci. Comput. 31 (6)
(2010) 4633–4653.
[13] M. Herty, Modeling, simulation and optimization of gas networks with compressors, Netw. Heterog. Media 2 (2007) 81–97.
[14] M. Herty, M. Seaïd, Simulation of transient gas flow at pipe-to-pipe intersections, J. Numer. Methods Fluids 56 (2008) 485–506.
[15] S. Jin, Z. Xin, The relaxation schemes for systems of conservation laws in arbitrary space dimensions, Commun. Pure Appl. Math. 48 (1995) 235–276.
[16] C.A. Dorao, M. Fernandino, Simulation of transients in natural gas pipelines, J. Natural Gas Sci. Eng. 3 (2011) 349–355.
[17] J. Brower, I. Gasser, M. Herty, Gas pipeline models revisited: Model hierarchies, nonisothermal models, and simulations of networks, Multiscale Model.
Simul. 9 (2011) 601–623.
[18] H. Colombo, R.M., G. Guerra, V. Schleper, Optimal control in networks of pipes and canals, SIAM J. Control Optim. 48 (3) (2009) 2032–2050.
[19] M. Chaczykowski, Sensitivity of pipeline gas flow model to the selection of the equation of state, Chem. Eng. Res. Des. 87 (2009) 1596–1603.
[20] A. Osiadacz, M. Chaczykowski, Verification of transient gas flow simulation model, in: Proceedings of the Conference on PSIG, PSIG, 2010.
[21] C.W. Gear, Numerical Initial Value Problems in Ordinary Differential Equations, Prentice-Hall, 1971.
[22] O. Bratland, Pipe flow 1. Single-phase flow assurance, 2009. ISBN 978-616-335-925-4.
[23] J.L. Modisette, J.P. Modisette, Transient and succession of steady-state pipeline flow models, in: Proceedings of the Conference on PSIG, PSIG, 2001.
[24] S.K. Bachman, M.J. Goodreau, Steady state - is the solution realistic for the piping network? in: Proceedings of the Conference on PSIG, PSIG, 2003.
[25] F. White, Fluid Mechanics, McGraw-Hill, 2002.
[26] J.H. Ferciger, M. Peric, Computational Methods for Fluid Dynamics, third ed., Springer, 2002.
[27] J.P. Modisette, Pipeline thermal models, in: Proceedings of the Conference on PSIG, PSIG, 2002.
[28] J.P. Modisette, Physics of pipeline flow, in: Proceedings of the Conference on PSIG, PSIG, 2003.
[29] U. Grigull, H. Sandner, Wärmeleitung, Springer-Verlag, 1986.
[30] J.H. Neher, The temperature rise of buried cables and pipes, Trans. Am. Inst. Electr. Eng. 68 (1) (1949) 9–21.
A. López-Benito et al. / Applied Mathematical Modelling 40 (2016) 10020–10037 10037

[31] J. Nikuradse, Laws of Flow in Rough Pipes (Translation of Strömungsgesetze in rauhen Rohren. VDI-Forschungsheft 361), National Advisory Committee
for Aeronautics (NACA), 1933.
[32] C.F. Colebrook, Turbulent flow in pipes, with particular reference to the transition regime between smooth and rough pipe laws, Inst. Civil Eng. J. 11
(4) (1939) 133–156.
[33] A. Uhl, Steady Flow in Gas Pipelines., Technical Report No. 10, Institute of Gas Technology, American Gas Association, 1965.
[34] N. Chen, An explicit equation for friction factor in pipe, Ind. Eng. Chem. Fundam. 1 (1979) 296–297.
[35] S. Haaland, Simple and explicit formulas for the friction factor in pipe flow, Inst. Civil Eng. J. 105 (1983) 89–90.
[36] K. Gersten, H.D. Papenfuss, T. Kurschat, P. Genillon, F.F. Pérez, N. Revell, New transmission-factor formula proposed for gas pipelines, Oil Gas J. 98 (7)
(20 0 0) 58–62.
[37] B.E. Poling, J.M. Prausnitz, J.P. O’Connel, The Properties of Gas and Liquids, fifth ed., McGraw-Hill, 2001.
[38] International Organization for Standardization, ISO-20765-1. Natural gas. Calculation of thermodynamic properties. Part 1: Gas phase properties for
transmission and distribution applications (2005).
[39] M. Cruzeix, A.L. Mignot, Analyse Numérique des Équations Différentielles, second ed., Masson, 1992.
[40] L.F. Shampine, Numerical Solution of Ordinary Differential Equations, Chapman & Hall, 1994.
[41] F.J. Elorza Tenreiro, A. López Benito, PipeCalc Project, Final Report, UPM-Enagas, 2004. (Internal report)
[42] F. Abdolahi, A. Mesbah, R.B. Boozarjomehry, W.Y. Svrcek, The effect of major parameters on simulation results of gas pipelines, Int. J. Mech. Sci. 49
(20 07) 989–10 0 0.
[43] The American Gas Asociation, Distribution-System Design, AGA, Arlington, 1990.

Вам также может понравиться