Open access
Technical Papers
Jun 3, 2021

Pressure Flow–Based Algorithms for Pressure-Driven Analysis of Water Distribution Networks

Publication: Journal of Water Resources Planning and Management
Volume 147, Issue 8

Abstract

In recent years, pressure-driven analyses of water distribution systems using pressure-flow relationships (PFRs) were shown to be more computationally efficient than those based on flow-pressure relationships (FPRs), such as the widely used Wagner equation and various spline approximations. Using a PFR enhances the convergence properties of the Newton–Raphson method used by most water distribution network (WDN) solvers. This paper derives a new way of incorporating a PFR into the classical Todini and Pilati global gradient algorithm (GGA). The convergence properties of the resulting solution algorithm are compared with those from two other existing PFR algorithms, EPANET 2.2 and the active-set method (ASM), as well as from a conventional flow-pressure–based algorithm on a number of different networks of varying size.

Introduction

Solvers for water distribution network (WDN) models, from the original Cross (1936) method to the widely used Todini and Pilati global gradient algorithm (GGA) (Todini and Pilati 1988), were essentially developed for design purposes and run by fixing demands while modifying pipe diameters until the desired pressures were reached at all network nodes. This implied using demand-driven approaches (DDAs) regardless of the resulting nodal pressures. With the advent of reliable codes, such as for instance EPANET (Rossman 2000) and several other research and commercial packages, the interest shifted from mere design to resilience and reliability-based design as well as to extended period simulations (EPS). As already mentioned, the design of WDNs was historically approached using the DDA where pipe diameters were to be found in order to meet desired pressures at nodes for a given topology and demands pattern. If pressures at all nodes were not met, diameters were increased until the objective was reached. In recent approaches, design also takes into account resilience and reliability criteria to allow overcoming critical situations and failures; moreover, leakage control can be improved by simulating in time using EPS leak losses that are usually represented as a function of the pipe average pressure (see for instance Giustolisi et al. 2008); finally, the increasingly extended availability of flow and pressure sensors, including smart meters, used in conjunction with data assimilation techniques (Hutton et al. 2014a, b; Bragalli et al. 2016) allows for offline as well as online EPS using WDN models aimed at increasing management capabilities while reducing risks of failures. Specifically, in EPS one does not change the pipe diameters, which means that pressure during the simulation in time, under certain conditions (Wu and Walski 2006) such as pipe bursts (or isolation) or during excessive water use, may fall below the desired pressure at some of the nodes, with the consequence that demand can’t be fully delivered. In these cases, DDAs fail to realistically represent the actual WDN behavior and, when pressure falls below the desired value, one needs to insert a pressure-based condition at the relevant nodes. This alternative pressure-driven approach (PDA) requires defining a flow-pressure relationship (FPR) to describe the reduced nodal flow delivered as a function of available pressure and demand.
Several analytical formulations of FPR were proposed in the literature (Bhave 1981; Germanopoulos 1985; Wagner et al. 1988; Reddy and Elango 1989, 1991; Chandapillai 1991; Fujiwara and Ganesharajah 1993; Tucciarelli et al. 1999; Tanyimboh et al. 2001; Wu et al. 2009; Tanyimboh and Templeman 2010; Jun and Guoping 2013; Morley and Tricarico 2014) while many other authors have introduced artificial components such as emitters, pipes, reservoirs, and so on, to represent the FPR behavior (Ang and Jowitt 2006; Baek et al. 2010; Suribabu and Neelakantan 2011; Jinesh Babu and Mohan 2012; Gorev and Kodzhespirova 2013; Sivakumar and Prasad 2014, 2015; Abdy Sayyed and Gupta 2013; Abdy Sayyed et al. 2014, 2015; Suribabu 2015; Mamizadeh and Shaoonizadeh 2016; Suribabu et al. 2017; Mahmoud et al. 2017). The interested reader can find a clear synthetic description of all the different approaches available in the literature in Suribabu et al. (2019).
Unfortunately, most of the FPRs were written in the form of demand d as a function of pressure head H, namely d=d(H), which, as discussed by Ackley et al. (2001) or by Todini (2006) and also recently demonstrated by Deuerlein et al. (2019), tends to increase the number of iterations required by a gradient approach, such as the Newton–Raphson (N–R), to reach convergence when solving the WDN problem.
The lack of convergence is mostly due to two reasons:
1.
The FPR is generally a concave function of H and also its first-order continuous derivatives approximations (Fujiwara and Ganesharajah 1993; Tanyimboh and Templeman 2010; Piller et al. 2003; Elhay et al. 2016; Deuerlein et al. 2019) unavoidably incorporate a concave portion instead of being a fully convex function as required by the N–R approach to rapidly converge;
2.
The first-order derivatives are noncontinuous for some suggested FPRs in the technical literature (e.g., Wagner et al. 1988) for H=Hm and H=H*, where H is the actual hydraulic head, H* is the desirable or nominal head, Hm the elevation at and below which no flow can be delivered, while Z is the geodetic elevation of the node.
To overcome these problems, several authors introduced underrelaxation coefficients or defined alternative first derivative continuous FPR, as subsequently discussed.
In 2020, the US Environmental Protection Agency (USEPA) released version 2.2 of EPANET (USEPA 2020) which uses an inverse formulation of the FPR by adding a virtual one-way link between a node and a virtual reservoir (fixed head node), to represent an inverse flow-pressure–type relation (FPR) in the form of H=H(d), which will be referred to as a pressure-flow relation (PFR).
This approach, already advocated and tested by Rossman several years before for describing emitters, resulted into a noticeable improvement in the convergence properties without increasing the size or the density of the system matrix. More recently, Deuerlein et al. (2019) found improved convergence performance using the inverse relations in their active-set method (ASM), the equations of which are derived by minimizing the Collins et al. (1978) “content” reaching the same system of equations used in EPANET 2.2, although solved differently, as subsequently discussed.
The aim of this work is to introduce a third derivation of the inverse FPR-based PDA algorithm in line with the original GGA derivation to point out the minor changes needed to convert the original demand-driven GGA to implement the PDA. The resulting equations are the same as the previous two derivations, but the final solution algorithm differs slightly from both.

GGA “Direct” Flow-Pressure d(H) Approach

The PDA problem is commonly solved in its “direct” FPR form to impose a reduction of the desired demand in the nodes where pressure is not sufficient to allow delivering it. From the classical GGA equations relevant to the DDA problem (Todini and Pilati 1988; Todini and Rossman 2013)
[A11A12A210][QH]=[A10H0d*]
(1)
where: QT=[Q1,Q2,,Qnp] unknown pipe flows (np is the number of WDN pipes); HT=[H1,H2,,Hnn] unknown nodal heads (nn is the number of WDN unknown head nodes); H0T=[Hnn+1,Hnn+2,,Hnt] known nodal heads (ntnn is the number of WDN nodes with known head); and d*T=[d1*,d2*,,dnn*] known nodal desired demands.
The matrices A12=A21T and A10 are incidence matrices defined as in Todini and Pilati (1988) with the flow considered positive if it is directed toward the relevant node and negative if it leaves the node, while A11 is a diagonal matrix, where the non-null elements are given by
A11(k,k)=rk|Qk|αk1
(2)
As proposed by Salgado-Castro (1988), Salgado-Castro et al. (1994), and subsequently by Todini (2006), one can derive the PDA equations by modifying the right-hand side vector to include the FPR equation. Without loss of generality, for instance, in this work we will use the FPR equation of Wagner et al. (1988) in the form
d(i)={0HiHimdi*(HiHim)1/c(Hi*Him)1/cHim<Hi<Hi*di*HiHi*
(3)
with c an exponent usually taken as c=2, to represent the unknown flow actually released at node i instead of the required di*.
Eq. (3) is graphically represented in Fig. 1, where Z is the geodetic elevation of the node, Him the height at and below which no flow can be delivered, and H* is the required service height at and above which full demand can be satisfied.
Fig. 1. Graphical representation of Wagner et al. (1988) FPR.
For a generic hypothetical solution Q and H, the system of Eq. (1) will not be satisfied with the equality sign. Therefore, we can write
[A11A12A210][QH]+[A10H0d]=[f1f2][00]
(4)
taking into account that d, given by Eq. (3), may be in part or entirely different from the desired demand d* being either null or a function of the hydraulic head H. Differentiating the system of Eq. (4), the result can be rearranged as
[D11A12A210][dQdH]+[0dH]=[D11A12A21D22][dQdH]=[df1df2]
(5)
where
D11(k,k)=αkrk|Qk|αk1
(6)
and
D22(i,i)=diHi={0HiHimdi*c(HiHim)1/c1(Hi*Him)1/cHim<Hi<Hi*0HiHi*
(7)
together with
df1=A11Q+A12H+A10H0df2=A21Qd
(8)
Taking into account the sign of the N–R algorithm, we will define all throughout this paper dQ=QQ(+) and dH=HH(+), leading to the following recursive solution provided by the GGA:
H(+)=A1FQ(+)=QD111(A11Q+A12H(+)+A10H0)
(9)
with H(+) and Q(+) the updated values of nodal hydraulic heads and pipe flows, and
A=A21D111A12D22
(10)
F=d+A21D111[(D11A11)QA10H0]D22H
(11)
The iterative process of Eq. (9) will proceed starting from an initial guess solution H1, and Q1.
The estimates of the actual delivered demands d(+) are then updated as a function of H(+) by means of Eq. (3). This solution only differs from the classical DDA solution for a different definition of dd* and the presence of the non-null diagonal matrix D22 which is subtracted from the main diagonal of the system and for a correction factor D22H applied to the right-hand side of the DDA problem.
Frequently, when using the direct FPR the convergence of the N–R process may be quite slow due to the concavity of the FPR which, similar to what happens with the N–R nodal approach (Shamir and Howard 1968), may lead to a rather slow alternating convergence (Todini and Pilati 1988; Todini and Rossman 2013).
To overcome this lack of convergence, various alternatives have been used, such as introducing underrelaxation coefficients (e.g., Giustolisi et al. 2008; Siew and Tanyimboh 2012; Elhay et al. 2016; Ciaponi and Creaco 2018). Other approaches, which assumed that the slow convergence was due to the lack of continuity of the first derivatives, defined alternative FPRs (Fujiwara and Ganesharajah 1993; Tanyimboh and Templeman 2010) and cubic functions or spline regularization (Piller et al. 2003; Elhay et al. 2016; Deuerlein et al. 2019) to guarantee the continuity of the first derivatives. Nonetheless, after a number of tests using these regularized functions, which still require underrelaxation to improve convergence, we reached the conclusion that more than the continuity of the first derivative, it is the presence of concavity that affects convergence and, in fact, most of the alternative FPRs introduced to guarantee continuous first-order derivatives are not strictly convex: indeed, they present an inflection point and the resulting functions are convex at one end but concave at the other end.

“Inverse” Pressure-Flow H(d) Approaches

EPANET 2.2 Method

In EPANET 2.2 (Rossman et al. 2020), the pressure-dependent demands are treated as unknown flows q=d in virtual one-way links connecting each node to a virtual reservoir with a fixed hydraulic head equal to the elevation at which the delivered flow becomes zero (Fig. 2). This is done by using an inverted Wagner et al. (1988) PFR in the interval 0<didi* with linear “barrier functions” added outside of this interval that impose a very large head loss for di0 or di>di*.
Fig. 2. Representing a pressure-dependent nodal demand with a virtual pipe and reservoir.
The idea of using an inverted form of a PFR in the GGA solver originates from the way in which emitters were modeled as unconstrained pressure-dependent demands in version 2.0 of EPANET released back in 2000 (Rossman 2000). In 2007, Rossman also described how the concept could be extended to model-bounded PFRs as well (Rossman 2007).
The EPANET 2.2 PDA algorithm can be derived as follows. Starting from the basic GGA system of equations
[A11A12A210][QH]=[A10H0d*]
(12)
whose definition is the same as per Eq. (1), the introduction of a set of new virtual links requires the addition of a vector d of nn unknowns representing the nodal demand flows. This leads to the revised system of equations
[A110A120A33IA21I0][QdH]=[A10H0Hm0]
(13)
where
A33(i,i)={0di0(Hi*Him)di*c|d|ic10<didi*0di>di*
(14)
I is an identity matrix of size nn×nn, while the nn vector of nodal desired demands is now set equal to zero being replaced by d in the system of equations. Please note that not all the diagonal elements of matrix A33 will be non-null.
For a given value for unknowns Q,d,H at a generic iteration, the system of equations will not necessarily be satisfied, namely
[A110A120A33IA21I0][QdH]+[A10H0Hm0][000]
(15)
Linearizing the problem by differentiating Eq. (15), the following system is obtained:
[D110A120D33IA21I0][dQdddH]=[df1df2df3]
(16)
where D11 is defined as in Eq. (6) while the diagonal matrix D22 is defined as
D33(i,i)={0di0c(Hi*Him)di*c|d|ic10<didi*0di>di*
(17)
and with
[df1df2df3]=[A11Q+A12H+A10H0A33dH+HmA21Qd]
(18)
The system to be solved at each iteration is
[dQdqdH]=[D110A120D33IA21I0]1[df1df2df3]
(19)
Which when solved by partitioning leads to
H(+)=(A)1FQ(+)=QD111(A11Q+A12H(+)+A10H0)d(+)=dD331(A33dH(+)+Hm)
(20)
with
A=A21D111A12+D331
(21)
F=FD331[(D33A33)dHm]
(22)
and F being the classical GGA expression
F=d+A21D111[(D11A11)QA10H0]
(23)
Please note that this formulation does not increase the size of the A matrix used to solve for new heads: it remains at nn×nn as in the original GGA. However, at this point, a direct solution of Eq. (20) is not possible because the diagonal matrix D33 may contain several zero values in its principal diagonal as per its definition given in Eq. (17), which makes it not invertible and therefore impossible to evaluate A or F using Eqs. (21) and (22). To solve this problem, EPANET 2.2 uses a barrier-function technique that adds large weights to the main diagonal of matrix A33 which is now defined as
A^33(i,i)={BIGdi0(Hi*Zi)di*c|d|ic10<didi*BIGdi>di*
(24)
with BIG a large positive number, such as 108. This leads to
D^33(i,i)={BIGdi0c(Hi*Him)di*c|d|ic10<didi*BIGdi>di*
(25)
with the result that D^33 will now be invertible.
Having modified A33 into A^33 an additional correction must be introduced in the second of Eq. (18) to match the original equations, by modifying Hm into H^m defined as
H^m(i)={Himdi0Him0<didi*Hi*BIGdi*di>di*
(26)
The new equations to be solved are finally
H(+)=(A^)1F^Q(+)=QD111(A11Q+A12H(+)+A10H0)d(+)=dD^331(A^33dH(+)+H^m)
(27)
with
A^=A21D111A12+D^331
(28)
and
F^=FD^331[(D^33A^33)dH^m]
(29)
Empirical testing has shown that the PDA performance of EPANET 2.2 could be improved by reducing the magnitude of the change in di produced by Eq. (27) (with its sign preserved) to di*/2 whenever it exceeds di*. It is this slightly modified EPANET, referred to as EPANET 2.2.1, that was used in the examples subsequently presented in this paper.

Active-Set Method

In 2019, Deuerlein et al. presented an inverse FPR approach based on the minimization of the Collins et al. (1978) “content.” As opposed to the scalar form of EPANET 2.2, this approach introduces, in matrix form, three sets of new constraints describing the inverse FPR. The resulting system of equations, which includes the Lagrange multipliers λi relevant to the nodes for which di0 and μi relevant to the nodes for which didi*, is then simplified to lead to a set of equations, Eq. (14) in Deuerlein et al. (2019). The resolving algorithm, the ASM, is then described in the “iteration loop” section and in Eqs. (15)–(17) of the aforementioned work.
In the ASM, the decision on which nodes to use the PFR is not only based on demand as in EPANET 2.2, but also on the Lagrange multipliers, namely either when 0<di<di*, or when di=0 and λi<0 (corresponding to Hi>Him) or, additionally, when di=di* and μi<0 (corresponding to Hi<Hi*). Following the decision on which nodes are in the limits and which ones fall between the limits, all estimated demands smaller than zero are set to zero and all estimated demand higher than the desired demand are set at the desired demand value.
The ASM solution equations also differ slightly from those of EPANET 2.2. Although formally the same, the equations are directly solved in ASM [Eqs. (15)–(17) in Deuerlein et al. (2019), while they are slightly modified in order to use the barrier function approach in EPANET 2.2. Using the ASM approach, Deuerlein et al. (2019) were able to show substantial advantages for using the inverse instead of the direct FPR in PDA on simple cases as well as on a set of large WDNs [Table 6 in Deuerlein et al. (2019)]

GGA-PDA Method

The proposed PDA extension of GGA follows a line of continuity with original GGA formulation (Todini and Pilati 1988) as well as with the approach described in Todini (2006). As it will be shown, the final equations to be solved remain formally the same of EPANET and the ASM, mostly differing in terms of the controls used to decide which are the nodes to be set at the upper or lower limit as opposed to the ones requiring being treated using the FPR curve.
In the case of PDA, the original set of equations [Eq. (4)] is insufficient to fully describe the problem since a number of actual demands d* will not be fully satisfied. Nonetheless, the solution can be found by expanding the original mathematical problem given in Eq. (4) with the introduction of a new set of unknowns δ, subject to the inequality constraints δi0 and δidi*, representing the amount of water to be delivered at the nodes together with a new set of equations describing the Wagner et al. (1988) inverse FPR, linking the new unknowns δ to the relevant hydraulic head
Hi=Him+Hi*Himdi*c|δi|c1δi
(30)
To solve the problem, either one correctly formulates the full minimization problem as in Deuerlein et al. (2019) or, due to the convexity of the “content” function, extends the Theil and van de Panne (1960) quadratic programming approach to convex constrained minimization problems, as already mentioned in Todini (2006), by repeatedly solving a series of unconstrained problems eliminating at each iteration from the active set of unknowns the variables not satisfying the constraints and setting them at their upper or lower limit. In Todini (2006), it was proposed to initially solve the DDA case described by Eq. (4) and to proceed iteratively by checking and eliminating the unknowns every time a new solution is found. Although the scheme is convergent, it obviously requires more iterations to be performed than checking and eliminating the unknowns at each N–R iteration, as is subsequently proposed.
The system of equations resulting from the unconstrained minimization of Collins et al. (1978) “content,” which not necessarily guarantees the estimated demands δ to fall in the range 0δd* is
[A110A120A33A32A21A230][QδH]=[A10H0Hmd]
(31)
where matrices A12=A21T and A10 are the same incidence matrices of the original GGA development; A11 is the original np×np diagonal matrix, with np the number of pipes, defined as in Eq. (2); and A33 is a new diagonal nδ×nδ matrix, with nδ the number of nodes for which the actual demand is unknown, defined as
A33(j,j)=(Hi(j)*Hi(j)m)di(j)*cδjc1
(32)
with j1,nδ; i1,nn and where j indicates the position of the jth node in the vector of δ and i(j) indicates the corresponding position in the vector of unknown pressure nodes.
Matrix A32=A23T is an incidence topological nδ×nn matrix filled with zeroes except having a 1 at nodes where actual demand is unknown. If nδ=nn, then A32=A23T=I, with I the identity matrix. The vector d contains the demands of vector d* which are set at the upper limit with all other elements being set to zero, allowing one to define the actual estimated demand vector d^ as
d^=dA23δ
(33)
In order to select the nodes where demand is set at the limiting value, we refer to the concept of power (energy per time unit) available at a node. In practice, nodes where d^i0 and HiHim or nodes where the power d^iHidi*Hi*, are treated as demand driven nodes by setting the demand of the first group to zero and the demand of the second group equal to the desired demand. Formally, all these nodes remain treated as in the DDA using the same equations in the original GGA. On the contrary, all the remaining nodes will be considered with unknown demand to be estimated using the additional equations introduced in Eq. (31).
If, at the end of an iteration all the demands are below or equal to their lower limit or above or equal to their upper limit, no modifications to the original GGA algorithm are required. In this case the recursive solution is the same of the DDA given in Todini and Pilati (1988) or in Todini and Rossman (2013). Alternatively, if after an iteration a number nδ of nodes strictly fall within the lower and upper limit, the solution must be found using the full set of Eq. (31) where vector d appearing on the right-hand side is defined for i1,nn as
d(i)=di={di*d^iHidi*Hi*0allothernodes
(34)
For a given value of unknowns Q,δ,H at some intermediate iteration, the system of equations will not necessarily be satisfied, namely
[A110A120A33A32A21A230][QδH]+[A10H0A32Hmd][000]
(35)
Linearizing the problem by differentiating Eq. (36), the following system is obtained:
[D110A120D33A32A21A230][dQdδdH]=[df1df2df3]
(36)
where D11 is defined as in Eq. (6) and with the diagonal matrix D33 given as
D33(j,j)=c(Hi(j)*Hi(j)m)di(j)*cδjc1
(37)
and where
[df1df2df3]=[A11Q+A12H+A10H0A33δ+A32(H+Hm)A21Q+A23δd]
(38)
Please note that all the D33(j,j) values are non-null, since δj always differs from zero in the range Hi(j)m<Hi(j)<Hi(j)* within which it has been computed, implying that the inverse of D33 always exists.
The system to be solved at each iteration is
[dQdδdH]=[D110A120D33A32A21A230]1[df1df2df3]
(39)
It is possible to solve it through partitioning by writing
[[dQdδ]dH]=[DNNT0]1[[df1df2]df3]
(40)
where
D=[D1100D33]
(41)
and
N=[A12A32];NT=[A21A23]
(42)
Solving by partitioning, one obtains:
H(+)=(A)1FQ(+)=QD111(A11Q+A12H(+)+A10H0)δ(+)=δD331[A33δ+A32(H(+)Hm)]
(43)
where H(+), Q(+) and δ(+) represent the updated unknown values.
In Eq. (43) δ(+) represents the updated values of the estimated demand d^(+) for the nodes strictly falling within the lower and upper limits. The fully updated set of estimated demands becomes
d^(+)=dA23δ(+)
(44)
The system matrix A and the right-hand side F are defined as
A=A+A23D331A32
(45)
F=F+A23D331[(D33A33)δ+A32Hm]
(46)
with A=A21D111A12 and F=d+A21D111[(D11A11)QA10H0]. Please also note that A23D331A32 is a diagonal matrix, only acting on part of the main diagonal of A, without increasing the complexity of the original DDA problem.
Therefore, the problem is again fully solved by iterating two equations formally identical to the ones given by Eq. (9), and the proposed solution only differs from the classical DDA solution for the presence of the non-null diagonal matrix A23D331A32 which is added to part of the main diagonal of the system and for a correction factor A23D331[(D33A33)δ+A32Hm] applied to part of the right-hand side.
Finally, before starting a new N–R iteration and after deciding which nodes will be at the lower or upper limit, all d^i<0 are set d^i=0 and all d^i>di* are set to d^i=di*.
It is not difficult to show that the equations in Eq. (43) are exactly the same equations given as Eqs. (15)–(17) in Deuerlein et al. (2019) and that they lead to the same solution of Eq. (27) of the EPANET 2.2 algorithm. Thus, the basic differences in the three algorithms are essentially limited to the selection criteria used to set the variables at their lower, intermediate, or upper limit, which can be summarized as in Table 1.
Table 1. Selection criteria for variables at the lower, intermediate, or upper limit used by the three approaches
 EPANETActive-set methodGGA-PDA
Lower limitd^i0(di*=0)(di*>0)[(d^i<0)(d^i=0HiHim)](di*=0)(di*>0)(d^i0HiHim)
Upper limitd^i>di*(di*>0)[(d^i>di*)(d^i=di*HiHi*)](di*>0)(d^iHidi*Hi*)
Within limits0<d^idi*(di*>0){[(d^i>0)(d^ii<di*)](d^i=0Hi>Him)(d^i=di*Hi<Hi*)}(di*>0)(d^i>0Hi>Him)(d^iHi<di*Hi*)
The last row of Table 1, which looks more complex, doesn’t require evaluation since all nodes not set at the lower or at the upper limit will fall within the limits.
Analogous to ASM, the advantages of the proposed GGA-PDA approach are:
1.
The solution is general and can be used for a variety of different FPRs;
2.
Only nodes within the lower and the upper limit are treated in PDA mode, all the others remaining in DDA;
3.
The inverse FPR solution improves the convergence properties of the gradient with respect to the one based on the direct FPR;
4.
The inverse FPR formulation does not require underrelaxation to improve convergence; and
5.
There is no need for defining additional artificial components such as pipes, reservoirs, flow control valves, check valves, and so on, as proposed by several authors (Suribabu et al. 2019).

Comparison of Results over a Number of Examples

The proposed approach was tested and compared to ASM and EPANET 2.2.1 on three extremely simple examples, already used as testbeds for other approaches and described in the Appendix, as well as on two larger networks, that can be found in the literature, to test the validity of the approach on more complex networks.
The first network example is a simple no-loop 5-node and 4-pipe example introduced by Gupta and Bhave (1996) and already used by Cheung et al. (2005), Ang and Jowitt (2006), Morley and Tricarico (2014), and others, to analyze the behavior of pressure-driven approaches. The second network example is the two-loop network, with 5 nodes and 6 pipes, recently used by Deuerlein et al. (2019) to test their ASM approach. The third example is based on the classical Hanoi network, introduced by Fujiwara and Khang (1990). A detailed description of pipe length and diameters as well as of nodal elevation, desired demand and hydraulic head for the first three examples can be found in the Appendix.
The two additional tested networks were the Modena water distribution network (Artina et al. 2011) and the Balerma Irrigation Network (Reca and Martinez 2006). The description of both networks can be widely found in the literature but also on the University of Exeter Center for Water Systems web-site (http://emps.exeter.ac.uk/engineering/research/cws/resources/benchmarks/design-resiliance-pareto-fronts/large-problems/). The Modena WDN, originally presented by Artina et al. (2011) includes 4 reservoirs, 268 nodes, and 317 pipes. In order to test the PDD approaches, the desired demand was multiplied by 2 in order to generate unmet demand nodes. The Balerma Irrigation Network example originally presented by Reca and Martinez (2006) includes 4 reservoirs, 443 nodes, and 454 pipes; for testing the PDD approaches, the Hazen–Williams equations with C=130 were used instead of the Darcy–Weisbach equations with ε=0.0025.

Head Losses

In all the examples, head losses were expressed in terms of the Hazen–Williams equation ΔH=[(10.67|Q|0.852Q)/(C1.852D4.8704)]L with ΔH the head loss in [m], Q the discharge in [m3/s], D and L the pipe diameter and length in [m], and C=130 for the all the networks excepted the second one where a value of C=145 was used to approach the actual delivered demands, around 22%–33% of desired demand, as described in Deuerlein et al. (2019) who used a Darcy–Weisbach loss equation.

Pressure-Flow Relation

The PFR used was the inverse Wagner et al. (1988) relation of Eq. (30). The PFR was only linearized at the very lower end for 0<δε, with ε1 a positive number (we used ε=106) to avoid the noninvertibility of matrix D33, as also discussed in Deuerlein et al. (2019).

Initial Values

The initial values assumed for the unknowns were chosen to conform the ones proposed in Deuerlein et al. (2019), namely
Qj=Ωj/3di=di*/2Hi=Him+(Hi*Him)/4
(47)
with Ωj the wetted area of pipe j.

Stopping Criterion

The stopping criterion was based on the default one used in EPANET 2.2, namely j=1np|Qj(+)Qj|/j=1np|Qj|<0.001.

EPANET Version

The EPANET 2.2 solver used for the test examples was a slightly modified version of the publicly available one. As previously described, its only change is to halve the computed change in a demand within an N–R iteration if it otherwise exceeds the full desired demand. The source code for this modified EPANET, referred to as version 2.2.1, has been posted to a public repository.

Analysis of Results

All the different pressure-driven algorithms produce essentially the same hydraulic solutions within the limits of the accuracy chosen stopping criterion. Full results for the first three smaller networks are listed in Appendix (Tables A2, A4, A6, A7, A8, and A9). Results for the two larger networks can be provided in the form of an Excel file upon request.
The convergence results are summarized in Table 2 where one can compare the number of iterations needed to reach convergence by the classical demand driven GGA and the alternative PDA. Table 2 also provides an indication on the number of nodes resulting at their lower and upper limits and the number of nodes falling within the limits. The three desired pressure values, H*Hm=30,20,10m, were in the range of operational WDNs. The fourth desired pressure value H*Hm=0.1m is an abstraction, only used to test convergence of the algorithms under stressed numerical conditions.
Table 2. Results in terms of number of iterations for the different cases and used approaches
TestbedsNumber of iterationsNumber of nodes
GGA DDADirect FPRInverse FPRNd=0N0<d<d*Nd=d*
WDNH*Hm (m)GGA PDAEPANET 2.2.1ASMGGA PDA
Network 12024444040
Network 22046666040
Network 330312a6770301
2012a7770274
1012a7770256
0.1No convergence11121101516
Network 4304445402680
205555024721
105555021058
0.1No convergence911113743188
Network 530357662640611
2067664337426
10107887329179
0.1No convergence11121215765221
a
With underrelaxation; no convergence otherwise.
As can be observed, the classical direct FPR-based approach converges fast in most situations, but sometimes, as in the case of the Net. 3, doesn’t converge at all and requires underrelaxation coefficients. This lack of convergence is also highlighted when dealing with the smaller pressure range. On the contrary, all PFR-based approaches, the one used in EPANET 2.2.1, the ASM, and the proposed GGA-PDA formulation, have an almost identical behavior and always converge without requiring underrelaxation, even in the stressing cases of a reduced Him/Hi* range.

Conclusions

This work proposes an approach using an inverse FPR in conjunction with the GGA, which has the advantage of leaving in demand-driven mode all the nodes for which the power is equal or higher than the power required to deliver at the node the desired demand at the required head, as well as all the nodes for which the estimated operating pressure and demand result at the same time smaller than or equal to zero, thus failing to deliver water.
The derived solution equations, based on an extension of the DDA GGA, are practically identical to the ones used in EPANET 2.2, which were derived in analogy with the use of a virtual pipe and a virtual reservoir connected to each node. They are also identical to the equations used in the ASM approach, which were derived starting from the constrained minimization of the “content” function. The three solutions only differ in terms of the criteria used to decide, at each N–R iteration, which are the nodes to be considered at the lower or upper limit and which ones fall in between.
The paper also shows the modest modifications that need to be applied to a DDA solver that only affect the main diagonal and the right-hand side of the linear system to be solved at each iteration.
Although demonstrated for the Wagner et al. (1988) equation, the algorithm performs well for any inverse strictly concave PFR, representing hydraulic losses, with the curvature typically ranging from linear to quadratic. On the contrary, inverse PFRs derived from first-order derivative continuous FPR functions [such as the ones proposed by Fujiwara and Ganesharajah (1993), Tanyimboh and Templeman (2010), or the cubic or splines functions] do not bring substantial improvements because they remain both concave and convex also after the inversion.
As demonstrated using a number of test examples, the proposed approach practically shows the same results and convergence properties of ASM and EPANET 2.2.1 over a variety of WDN and a range of desired pressures.

Appendix. Description of Examples

Network 1

The first example is the simple line WDN, displayed in Fig. 3, introduced by Gupta and Bhave (1996) and also used by Cheung et al. (2005), Ang and Jowitt (2006), and Morley and Tricarico (2014).
Fig. 3. First network example.
The description of the network is given in Table 3 and the simulation results are summarized in Table 4, where the ratio d/d* represents the fulfillment rate of the water demand.
Table 3. Basic information for Network 1
PipeLength (m)Diameter (mm)C (T1.852L4.556)NodeZ (m)Hm (m)d* (m3/h)H* (m)
P11,000400130N1100
P21,000350130N29090120110
P31,000300130N38888120108
P41,000300130N49090180110
 N58585240105
Table 4. Network 1 solution
PipeDischarge (m3/h)Noded (m3/h)d/d* (%)H (m)
P1375.15N1100
P2297.89N277.266498.29
P3221.26N376.636496.16
P4145.46N475.804293.55
  N5145.466192.35

Network 2

The second example is the WDN, displayed in Fig. 4, introduced by Deuerlein et al. (2019) to demonstrate the convergence of their PDA based on the minimization of the “content” (Collins et al. 1978).
Fig. 4. Second network example.
The description of the network is given in Table 5 and the simulation results are summarized in Table 6.
Table 5. Basic information for Network 2
PipeLength (m)Diameter (mm)C (T1.852L4.556)NodeZ (m)Hm (m)d* (m3/h)H* (m)
P1500300145N120
P2500300145N20.00.090020
P3500300145N30.00.090020
P4500300145N40.00.01,35020
P5500300145N50.00.090020
P6500300145 
Table 6. Network 2 solution
PipeDischarge (m3/h)Noded (m3/h)d/d* (%)H (m)
P11,010.95N120
P2232.63N2300.74332.23
P3242.89N3207.58151.06
P4234.68N4296.86330.97
P525.06N5205.74231.05
P653.96 

Network 3

The third example known as Hanoi WDN, is displayed in Fig. 5, introduced by Fujiwara and Khang (1990) and widely used to test optimization approaches.
Fig. 5. Third network example: the Hanoi WDN.
The description of the network is given in Table 7, while the simulation results are summarized in Table 8 (H*Hm=30m), Table 9 (H*Hm=20m), Table 10 (H*Hm=10m), and Table 11 (H*Hm=0.1m).
Table 7. Basic information for Network 3
PipeLength (m)Diameter (mm)C (T1.852L4.556)NodeZ (m)Hm (m)d* (m3/h)H* (m)
P1100800130N1100
P21,350800130N201089040302010.1
P3900800130N301085040302010.1
P41,150800130N401013040302010.1
P51,450800130N501072540302010.1
P6450800130N60101,00540302010.1
P7850800130N70101,35040302010.1
P8850800130N801055040302010.1
P9800800130N901052540302010.1
P10950800130N1001052540302010.1
P111,200800130N1101050040302010.1
P123,500800130N1201056040302010.1
P13800800130N1301094040302010.1
P14500800130N1401061540302010.1
P15550800130N1501028040302010.1
P162,730800130N1601031040302010.1
P171,750800130N1701086540302010.1
P18800800130N180101,34540302010.1
P19400800130N190106040302010.1
P202,200800130N200101,27540302010.1
P211,500800130N2101093040302010.1
P22500800130N2201048540302010.1
P232,650800130N230101,04540302010.1
P241,230800130N2401082040302010.1
P251,300800130N2501017040302010.1
P26850800130N2601090040302010.1
P27300800130N2701037040302010.1
P28750800130N2801029040302010.1
P291,500800130N2901036040302010.1
P302,000800130N3001036040302010.1
P311,600800130N3101010540302010.1
P32150800130N3201080540302010.1
P33860800130 
P34950800130
Table 8. Network 3 solution for H*Hm=30  m
H*=40  m
PipeDischarge (m3/h)Noded (m3/h)d/d* (%)H (m)
P114,061.79N1100
P213,171.79N2890.0010095.20
P33,936.79N3818.119637.79
P43,821.24N4115.558933.70
P53,247.92N5573.327928.76
P62,557.76N6690.166924.15
P71,661.32N7896.446623.23
P81,307.06N8354.276422.45
P9975.77N9331.296321.95
P101,196.38N10327.466221.67
P11890.93N11305.456121.20
P12554.18N12336.756020.85
P13548.07N13554.185920.43
P14933.22N14385.146321.77
P151,109.74N15176.526321.92
P162,350.80N16197.396422.16
P173,000.71N17649.917526.94
P184,145.75N181,145.038531.74
P194,201.32N1955.589335.74
P204,215.57N20944.097426.45
P211,029.21N21676.637325.88
P22352.58N22352.587325.85
P232,242.27N23666.526422.20
P24887.34N24515.376321.85
P25371.97N25106.516321.78
P26246.51N26564.416321.80
P27810.92N27232.756321.87
P281,043.68N28182.916321.93
P29688.40N29225.126321.73
P30505.49N30224.606221.68
P31280.36N3165.516221.68
P3255.76N32502.236221.68
P339.74 
P34511.97
Table 9. Network 3 solution for H*Hm=20  m
H*=30  m
PipeDischarge (m3/h)Noded (m3/h)d/d* (%)H (m)
P114,459.62N1100
P213,569.62N2890.0010094.95
P34,036.48N3850.0010034.28
P43,906.48N4130.0010030.00
P53,281.70N5624.798624.85
P62,565.69N6716.017120.15
P71,648.76N7916.936819.23
P81,291.12N8357.636518.46
P9959.78N9331.346317.97
P101,166.97N10325.766217.70
P11866.01N11300.976017.25
P12536.69N12329.325916.92
P13532.95N13536.695716.52
P14916.77N14383.826217.79
P151,093.22N15176.456317.94
P162,338.28N16198.206418.18
P173,033.00N17694.728022.90
P184,302.03N181,269.039427.80
P194,362.03N1960.0010032.09
P204,321.11N201,004.347922.41
P211,085.64N21713.837721.78
P22371.81N22371.817721.75
P232,231.14N23669.326418.20
P24878.82N24513.966317.86
P25364.85N25106.066217.79
P26252.21N26562.376217.81
P27814.57N27232.286317.88
P281,046.86N28182.716317.94
P29683.00N29223.956217.74
P30500.29N30223.186217.69
P31276.34N3165.096217.69
P3253.15N32499.066217.69
P3311.94 
P34511.00
Table 10. Network 3 solution for H*Hm=10  m
H*=20  m
PipeDischarge (m3/h)Noded (m3/h)d/d* (%)H (m)
P114,868.69N1100
P213,978.69N2890.0010094.68
P34,152.97N3850.0010030.59
P44,022.97N4130.0010026.07
P53,297.97N5725.007720.64
P62,526.64N6771.337115.89
P71,572.87N7953.776514.99
P81,212.82N8360.056214.29
P9887.1N9325.726013.85
P101,070.84N10315.845713.62
P11786.59N11284.265413.23
P12482.12N12304.465112.96
P13499.58N13482.126112.63
P14873.6N14374.026213.70
P151,047.08N15173.476413.84
P162,294.19N16197.379314.05
P173,097.06N17802.8610018.61
P184,442.06N181,345.0010023.71
P194,502.06N1960.008928.26
P204,473.66N201,134.948517.92
P211,197.35N21787.608417.17
P22409.75N22409.756317.14
P232,141.36N23663.116114.03
P24828.44N24499.796013.71
P25328.65N25102.786113.66
P26276.66N26546.256113.68
P27822.9N27226.846213.76
P281,049.75N28178.406013.78
P29649.82N29216.186013.61
P30471.42N30214.806013.56
P31255.25N3162.656013.56
P3240.45N32480.3210013.56
P3322.2 
P34502.52
Table 11. Network 3 solution for H*Hm=0.1  m
H*=10.1  m
PipeDischarge (m3/h)Noded (m3/h)d/d* (%)H (m)
P115,244.69N1100
P214,354.69N2890.0010094.43
P34,299.68N3850.0010027.10
P44,169.68N4130.0010022.29
P53,444.68N5725.0010016.48
P62,439.68N61,005.0010011.34
P71,089.68N71,350.0010010.50
P8539.68N8550.0010010.14
P9205.27N9334.406410.04
P10211.72N10263.955010.03
P1188.51N11123.212510.01
P1223.72N1264.791210.00
P13270.4N1323.72310.00
P14708.55N14438.157110.05
P15988.55N15280.0010010.15
P162,280.09N16310.0010010.34
P173,145.09N17865.0010014.85
P184,490.09N181,345.0010020.09
P194,550.09N1960.0010024.73
P204,654.92N201,275.0010013.48
P211,415N21930.0010012.45
P22485N22485.0010012.40
P231,964.92N231,045.0010010.15
P24510.24N24413.845010.03
P2596.4N2574.764410.02
P26161.86N26491.355510.03
P27653.21N27328.338910.08
P28981.55N28203.677010.05
P29409.68N29118.473310.01
P30206.01N3076.512110.00
P3187.55N3122.282110.00
P3211.04N32172.262110.00
P3311.24 
P34183.51

Network 4

The Modena WDN is shown in Fig. 6, originally presented by Artina et al. (2011). The network includes 268 nodes and 317 pipes and it is supplied by four reservoirs.
Fig. 6. Fourth network example: the Modena WDN.

Network 5

The Balerma Irrigation Network shown in Fig. 7, first presented by Reca and Martinez (2006), is an adaption of an existing irrigation network in the Sol-Poniente irrigation district, located in Balerma in the province of Almería (Spain). The network includes 443 nodes and 454 pipes and it is supplied by four reservoirs.
Fig. 7. Fifth network example: the Balerma Irrigation Network.
Description of the Balerma Irrigation Network and its relevant data can be found at http://emps.exeter.ac.uk/engineering/research/cws/resources/benchmarks/design-resiliance-pareto-fronts/large-problems/.

Data Availability Statement

The EPANET 2.2.1 input files for the test cases, together with the APLX codes and input data for both the GGA-PDA and ASM approaches are available from the corresponding author upon reasonable request. The source code for the modified EPANET 2.2.1 used in this study can be found at https://github.com/OpenWaterAnalytics/EPANET/tree/dev-PDA_mod.

Acknowledgments

The authors would like to express their gratitude to the reviewers, who dedicated a considerable amount of time in the revision of this manuscript, which was greatly improved after their comments and suggestions. Support for this paper was provided in part by the research project: SCN_00489 SMART WATERTECH: Smart Community for the Development and Application of Monitoring Technologies and Innovative Control Systems for an Integrated Water Service (Ministero dell’Università e della Ricerca–Italian Government).

References

Abdy Sayyed, M. A. H., and R. Gupta. 2013. “Predicting deficient condition performance of water distribution networks.” J. Civ. Eng. Infrastruct. 46 (2): 161–173. https://doi.org/10.7508/ceij.2013.02.004.
Abdy Sayyed, M. A. H., R. Gupta, and T. T. Tanyimboh. 2014. “Modelling pressure deficient water distribution networks in EPANET.” Procedia Eng. 89: 626–631. https://doi.org/10.1016/j.proeng.2014.11.487.
Abdy Sayyed, M. A. H., R. Gupta, and T. T. Tanyimboh. 2015. “Noniterative application of EPANET for pressure dependent modelling of water distribution systems.” Water Resour. Manage. 29 (9): 3227–3242. https://doi.org/10.1007/s11269-015-0992-0.
Ackley, J. R. L., T. T. Tanyimboh, B. Tahar, and A. B. Templeman. 2001. “Head driven analysis of water distribution systems.” In Vol. 1 of Water software systems: Theory and applications, edited by B. Ulanicki, B. Coulbeck, and J. P. Rance, 183–192. Baldock, UK: Research Studies Press.
Ang, W. K., and P. W. Jowitt. 2006. “Solution for water distribution systems under pressure-deficient conditions.” J. Water Res. Plann. Manage. 132 (3): 175–182. https://doi.org/10.1061/(ASCE)0733-9496(2006)132:3(175).
Artina, S., A. Bolognesi, C. Bragalli, C. D’Ambrosio, and A. Marchi. 2011. “Comparison among best solutions of the optimal design of water distribution networks obtained with different algorithms.” In Vol. 3 of Proc., CCWI 2011 Urban Water Management: Challenges and Opportunities, 985–990. Exeter, UK: Centre for Water Systems.
Baek, C. W., H. D. Jun, and J. H. Kim. 2010. “Development of a PDA model for water distribution systems using harmony search algorithm.” KSCE J. Civ. Eng. 14 (4): 613–625. https://doi.org/10.1007/s12205-010-0613-7.
Bhave, P. R. 1981. “Node flow analysis of water distribution systems.” J. Transp. Eng. 107 (4): 457–467. https://doi.org/10.1061/TPEJAN.0000938.
Bragalli, C., M. Fortini, and E. Todini. 2016. “Enhancing knowledge in water distribution systems via data assimilation.” Water Resour. Manage. 30 (11): 3689–3706. https://doi.org/10.1007/s11269-016-1372-0.
Chandapillai, J. 1991. “Realistic simulation of water distribution system.” J. Transp. Eng. 117 (2): 258–263. https://doi.org/10.1061/(ASCE)0733-947X(1991)117:2(258).
Cheung, P. B., J. E. Van Zyl, and L. F. R. Reis. 2005. “Extension of EPANET for pressure driven demand modeling in water distribution system.” Vol. 1 of Computing and Control for the Water Industry\, 311–316. Exeter, UK: Centre for Water Systems.
Ciaponi, C., and E. Creaco. 2018. “Comparison of pressure-driven formulations for WDN simulation.” Water 10 (4): 523. https://doi.org/10.3390/w10040523.
Collins, M., L. Cooper, R. Helgason, J. Kenningston, and L. Le Blanc. 1978. “Solving the pipe network analysis problem using optimization techniques.” Manage. Sci. 24 (7): 747–760. https://doi.org/10.1287/mnsc.24.7.747.
Cross, H. 1936. Analysis of flow in networks of conduits or conductors. Urbana, IL: Engineering Experiment Station, Univ. of Illinois.
Deuerlein, J. W., O. Piller, S. Elhay, and A. R. Simpson. 2019. “Content-based active-set method for the pressure-dependent model of water distribution systems.” J. Water Resour. Plann. Manage. 145 (1): 04018082. https://doi.org/10.1061/(ASCE)WR.1943-5452.0001003.
Elhay, S., O. Piller, J. Deuerlein, and A. Simpson. 2016. “A robust, rapidly convergent method that solves the water distribution equations for pressure-dependent models.” J. Water Resour. Plann. Manage. 142 (2): 04015047. https://doi.org/10.1061/(ASCE)WR.1943-5452.0000578.
Fujiwara, O., and T. Ganesharajah. 1993. “Reliability assessment of water supply systems with storage and distribution networks.” Water Resour. Res. 29 (8): 2917–2924. https://doi.org/10.1029/93WR00857.
Fujiwara, O., and D. B. Khang. 1990. “A two-phase decomposition method for optimal design of looped water distribution networks.” Water Resour. Res. 26 (4): 539–549. https://doi.org/10.1029/WR026i004p00539.
Germanopoulos, G. 1985. “A technical note on the inclusion of pressure-dependent demand and leakage terms in water supply network models.” Civ. Eng. Syst. 2 (3): 171–179. https://doi.org/10.1080/02630258508970401.
Giustolisi, O., D. Savic, and Z. Kapelan. 2008. “Pressure-driven demand and leakage simulation for water distribution networks.” J. Hydraul. Eng. 134 (5): 626–635. https://doi.org/10.1061/(ASCE)0733-9429(2008)134:5(626).
Gorev, N. B., and I. F. Kodzhespirova. 2013. “Noniterative implementation of pressure-dependent demands using the hydraulic analysis engine of EPANET 2.” Water Resour. Manage. 27 (10): 3623. https://doi.org/10.1007/s11269-013-0369-1.
Gupta, R., and P. R. Bhave. 1996. “Comparison of methods for predicting deficient-network performance.” J. Water Resour. Plann. Manage. 122 (3): 214–217. https://doi.org/10.1061/(ASCE)0733-9496(1996)122:3(214).
Hutton, C., Z. S. Kapelan, L. Vamvakeridou-Lyroudia, and D. Savić. 2014b. “Dealing with uncertainty in water distribution system models: A framework for real-time modeling and data assimilation.” J. Water Resour. Plann. Manage. 140 (2): 169–183. https://doi.org/10.1061/(ASCE)WR.1943-5452.0000325.
Hutton, C. J., Z. S. Kapelan, L. Vamvakeridou-Lyroudia, and D. Savić. 2014a. “Application of formal and informal Bayesian methods for water distribution hydraulic model calibration.” J. Water Resour. Plann. Manage. 140 (11): 04014030. https://doi.org/10.1061/(ASCE)WR.1943-5452.0000412.
Jinesh Babu, K. S., and S. Mohan. 2012. “Extended period simulation for pressure-deficient water distribution network.” J. Comput. Civ. Eng. 26 (4): 498–505. https://doi.org/10.1061/(ASCE)CP.1943-5487.0000160.
Jun, L., and Y. Guoping. 2013. “Iterative methodology of pressure de- pendent demand based on EPANET for pressure-deficient water distribution analysis.” J. Water Res. Plann. Manage. 139 (1): 34–44. https://doi.org/10.1061/(ASCE)WR.1943-5452.0000227.
Mahmoud, H. A., D. Aavic, and Z. Kapelan. 2017. “New pressure-driven approach for modeling water distribution networks.” J. Water Res. Plann. Manage. 143 (8): 04017031. https://doi.org/10.1061/(ASCE)WR.1943-5452.0000781.
Mamizadeh, J., and S. Shaoonizadeh. 2016. “Application of modified complementary reservoir approach in analysis of water distribution networks under pressure deficient conditions.” Urban Water J. 14 (4): 386–393. https://doi.org/10.1080/1573062X.2016.1171884.
Morley, M. S., and C. Tricarico. 2014. Pressure driven demand extension for EPANET (EPANETpdd). Exeter, UK: Univ. of Exeter.
Piller, O., B. Bremond, and M. Poulton. 2003 “Least action principles appropriate to pressure driven models of pipe networks.” In Proc., World Water & Environmental Resources Congress 2003, 1–15. Reston, VA: ASCE. https://doi.org/10.1061/40685(2003)113.
Reca, J., and J. Martinez. 2006. “Genetic algorithms for the design of looped irrigation water distribution networks.” Water Resour. Res. 42 (5): 1–9. https://doi.org/10.1029/2005WR004383.
Reddy, L. S., and K. Elango. 1989. “Analysis of water distribution networks with head dependent outlets.” Civ. Eng. Syst. 6 (3): 102–110. https://doi.org/10.1080/02630258908970550.
Reddy, L. S., and K. Elango. 1991. “A new approach to the analysis of water starved networks.” J. Indian Water Works Assoc. 23 (1): 31–38.
Rossman, L. A. 2000. EPANET 2 users’ manual. EPA/600/R-00/057. Cincinnati: National Risk Management Research Laboratory, USEPA.
Rossman, L. A. 2007. “Discussion of ‘Solution for water distribution systems under pressure-deficient conditions’ by Wah Khim Ang and Paul W. Jowitt.” J. Water Res. Plann. Manage. 133 (6): 566–567. https://doi.org/10.1061/(ASCE)0733-9496(2007)133:6(566.2).
Rossman, L. A., H. Woo, M. Tryby, F. Shang, R. Janke, and T. Haxton. 2020. EPANET 2.2 user’s manual. Cincinnati: Water Infrastructure Division, Center for Environmental Solutions and Emergency Response, Office of Research and Development, USEPA.
Salgado-Castro, R. O. 1988. “Computer modeling of water supply distribution network using the gradient method.” Ph.D. thesis, Dept. of Civil Engineering, Univ. of Newcastle-Upon-Tyne.
Salgado-Castro, R. O., J. Rojo, and S. Zepeda. 1994. “Extended gradient method for fully non-linear head and flow analysis in pipe networks.” In Vol. 1 of Integrated computer applications in water supply, edited by B. Coulbeck, 49–60. New York: Wiley.
Shamir, U., and C. D. D. Howard. 1968. “Water distribution systems analysis.” J. Hydraul. Div. 94 (1): 219–234. https://doi.org/10.1061/JYCEAJ.0001747.
Siew, C., and T. T. Tanyimboh. 2012. “Pressure-dependent EPANET extension.” Water Resour. Manage. 26 (6): 1477–1498. https://doi.org/10.1007/s11269-011-9968-x.
Sivakumar, P., and R. K. Prasad. 2014. “Simulation of water distribution network under pressure-deficient condition.” Water Resour. Manage. 28 (10): 3271–3290. https://doi.org/10.1007/s11269-014-0677-0.
Sivakumar, P., and R. K. Prasad. 2015. “Extended period simulation of pressure-deficient networks using pressure reducing valves.” Water Resour. Manage. 29 (5): 1713–1730. https://doi.org/10.1007/s11269-014-0907-5.
Suribabu, C. R. 2015. “Emitter based approach for estimation of nodal outflow to pressure deficient water distribution networks under pressure management.” Sci. Iran. 22 (5): 1765–1778.
Suribabu, C. R., and T. R. Neelakantan. 2011. “Balancing reservoir based approach for solution to pressure deficient water distribution networks.” Int. J. Civ. Struct. Eng. 2 (2): 639–647.
Suribabu, C. R., T. R. Neelakantan, and P. Sivakumar. 2017. “Improved complementary reservoir solution to evaluate nodal outflow under pressure deficient conditions.” ISH J. Hydraul. Eng. 23 (3): 260–266. https://doi.org/10.1080/09715010.2017.1298060.
Suribabu, C. R., N. T. Renganathan, S. Perumal, and D. Paez. 2019. “Analysis of water distribution network under pressure-deficient conditions through emitter setting.” Drink. Water Eng. Sci. 12 (1): 1–13. https://doi.org/10.5194/dwes-12-1-2019.
Tanyimboh, T. T., M. Tabesh, and R. Burrows. 2001. “Appraisal of source head methods for calculating reliability of water distribution networks.” J. Water Res. Plann. Manage. 127 (4): 206–213. https://doi.org/10.1061/(ASCE)0733-9496(2001)127:4(206).
Tanyimboh, T. T., and A. B. Templeman. 2010. “Seamless pressure-deficient water distribution system model.” Proc. Inst. Civ. Eng. Water Manage. 163 (8): 389–396. https://doi.org/10.1680/wama.900013.
Theil, H., and C. van de Panne. 1960. “Quadratic programming as an extension of conventional quadratic maximization.” Manage. Sci. 7 (1): 1–20. https://doi.org/10.1287/mnsc.7.1.1.
Todini, E. 2006. “Towards realistic extended period simulations (EPS) in looped pipe networks.” In Proc., 8th Annual Water Distribution Systems Analysis Symp., 1–16. Reston, VA: ASCE. https://doi.org/10.1061/9780784409411.
Todini, E., and S. Pilati. 1988. “A gradient method for the solution of looped pipe networks.” In Vol. 1 of Computer applications in water supply: System analysis and simulation, edited by B. Coulbeck and C. H. Orr, 1–20. London: Wiley.
Todini, E., and L. Rossman. 2013. “Unified framework for deriving simultaneous equation algorithms for water distribution networks.” J. Hydraul. Eng. 139 (5): 511–526. https://doi.org/10.1061/(ASCE)HY.1943-7900.0000703.
Tucciarelli, T., A. Criminisi, and D. Termini. 1999. “Leak analysis in pipeline systems by means of optimal valve regulation.” J. Hydraul. Eng. 125 (3): 277–285. https://doi.org/10.1061/(ASCE)0733-9429(1999)125:3(277).
USEPA. 2020. “EPANET application for modeling drinking water distribution systems.” Accessed May 14, 2021. https://www.epa.gov/water-research/epanet.
Wagner, J. M., U. Shamir, and D. H. Marks. 1988. “Water distribution reliability: Simulation methods.” J. Water Res. Plann. Manage. 114 (3): 276–294. https://doi.org/10.1061/(ASCE)0733-9496(1988)114:3(276).
Wu, Z. Y., and T. Walski. 2006. “Pressure dependent hydraulic modelling for water distribution systems under abnormal conditions.” In Proc., IWA World Water Congress and Exhibition. Beijing: IWA Publishing.
Wu, Z. Y., R. H. Wang, T. M. Walski, S. Y. Yang, D. Bowdler, and C. C. Baggett. 2009. “Extended global-gradient algorithm for pressure-dependent water distribution analysis.” J. Water Res. Plann. Manage. 135 (1): 13–22. https://doi.org/10.1061/(ASCE)0733-9496(2009)135:1(13).

Information & Authors

Information

Published In

Go to Journal of Water Resources Planning and Management
Journal of Water Resources Planning and Management
Volume 147Issue 8August 2021

History

Received: Dec 17, 2019
Accepted: Feb 1, 2021
Published online: Jun 3, 2021
Published in print: Aug 1, 2021
Discussion open until: Nov 3, 2021

Authors

Affiliations

Retired Professor, Honorary President of Italian Hydrological Society, Piazza di Porta S. Donato 1, Bologna 40127, Italy (corresponding author). ORCID: https://orcid.org/0000-0003-0847-1379. Email: [email protected]
Simone Santopietro [email protected]
Researcher, Dipartimento di Ingegneria Civile e Meccanica, Univ. of Cassino and Southern Lazio, Via G. Di Biasio 43, Cassino (FR) 03043, Italy. Email: [email protected]
Rudy Gargano [email protected]
Associate Professor, Dipartimento di Ingegneria Civile e Meccanica, Univ. of Cassino and Southern Lazio, Via G. Di Biasio 43, Cassino (FR) 03043, Italy. Email: [email protected]
Lewis A. Rossman, M.ASCE [email protected]
Consulting Engineer, Cincinnati, OH 45224. Email: [email protected]

Metrics & Citations

Metrics

Citations

Download citation

If you have the appropriate software installed, you can download article citation data to the citation manager of your choice. Simply select your manager software from the list below and click Download.

Cited by

  • Turbine modeling for steady‐state analysis in hydropower plant networks with complex layouts using a modified global gradient algorithm, Energy Science & Engineering, 10.1002/ese3.1390, 11, 3, (1251-1269), (2023).
  • Development and Testing of a Mathematical Model for Dynamic Network Simulation of the Oil Displacement Process, Fluids, 10.3390/fluids7090311, 7, 9, (311), (2022).
  • Comparing alternative PDA solvers with EPANET, Journal of Hydroinformatics, 10.2166/hydro.2022.005, 24, 3, (697-710), (2022).
  • High-Order Global Algorithm for the Pressure-Driven Modeling of Water Distribution Networks, Journal of Water Resources Planning and Management, 10.1061/(ASCE)WR.1943-5452.0001524, 148, 3, (2022).

View Options

Media

Figures

Other

Tables

Share

Share

Copy the content Link

Share with email

Email a colleague

Share