
In this paper, we consider a stationary model for the flow through a network. The flow is determined by the values at the boundary nodes of the network. We call these values the loads of the network. In the applications, the feasible loads must satisfy some box constraints. We analyze the structure of the set of feasible loads. Our analysis is motivated by gas pipeline flows, where the box constraints are pressure bounds.
We present sufficient conditions that imply that the feasible set is star-shaped with respect to special points. Under stronger conditions, we prove the convexity of the set of feasible loads. All the results are given for passive networks with and without compressor stations.
This analysis is motivated by the aim to use the spheric-radial decomposition for stochastic boundary data in this model. This paper can be used for simplifying the algorithmic use of the spheric-radial decomposition.
Citation: Martin Gugat, Rüdiger Schultz, Michael Schuster. Convexity and starshapedness of feasible sets in stationary flow networks[J]. Networks and Heterogeneous Media, 2020, 15(2): 171-195. doi: 10.3934/nhm.2020008
[1] | Martin Gugat, Rüdiger Schultz, Michael Schuster . Convexity and starshapedness of feasible sets in stationary flow networks. Networks and Heterogeneous Media, 2020, 15(2): 171-195. doi: 10.3934/nhm.2020008 |
[2] | Martin Gugat, Falk M. Hante, Markus Hirsch-Dick, Günter Leugering . Stationary states in gas networks. Networks and Heterogeneous Media, 2015, 10(2): 295-320. doi: 10.3934/nhm.2015.10.295 |
[3] | Mapundi K. Banda, Michael Herty, Axel Klar . Gas flow in pipeline networks. Networks and Heterogeneous Media, 2006, 1(1): 41-56. doi: 10.3934/nhm.2006.1.41 |
[4] | Michael Herty . Modeling, simulation and optimization of gas networks with compressors. Networks and Heterogeneous Media, 2007, 2(1): 81-97. doi: 10.3934/nhm.2007.2.81 |
[5] | Markus Dick, Martin Gugat, Günter Leugering . Classical solutions and feedback stabilization for the gas flow in a sequence of pipes. Networks and Heterogeneous Media, 2010, 5(4): 691-709. doi: 10.3934/nhm.2010.5.691 |
[6] | Mapundi K. Banda, Michael Herty, Axel Klar . Coupling conditions for gas networks governed by the isothermal Euler equations. Networks and Heterogeneous Media, 2006, 1(2): 295-314. doi: 10.3934/nhm.2006.1.295 |
[7] | Jens Lang, Pascal Mindt . Entropy-preserving coupling conditions for one-dimensional Euler systems at junctions. Networks and Heterogeneous Media, 2018, 13(1): 177-190. doi: 10.3934/nhm.2018008 |
[8] | Gen Qi Xu, Siu Pang Yung . Stability and Riesz basis property of a star-shaped network of Euler-Bernoulli beams with joint damping. Networks and Heterogeneous Media, 2008, 3(4): 723-747. doi: 10.3934/nhm.2008.3.723 |
[9] | Markus Musch, Ulrik Skre Fjordholm, Nils Henrik Risebro . Well-posedness theory for nonlinear scalar conservation laws on networks. Networks and Heterogeneous Media, 2022, 17(1): 101-128. doi: 10.3934/nhm.2021025 |
[10] | Gunhild A. Reigstad . Numerical network models and entropy principles for isothermal junction flow. Networks and Heterogeneous Media, 2014, 9(1): 65-95. doi: 10.3934/nhm.2014.9.65 |
In this paper, we consider a stationary model for the flow through a network. The flow is determined by the values at the boundary nodes of the network. We call these values the loads of the network. In the applications, the feasible loads must satisfy some box constraints. We analyze the structure of the set of feasible loads. Our analysis is motivated by gas pipeline flows, where the box constraints are pressure bounds.
We present sufficient conditions that imply that the feasible set is star-shaped with respect to special points. Under stronger conditions, we prove the convexity of the set of feasible loads. All the results are given for passive networks with and without compressor stations.
This analysis is motivated by the aim to use the spheric-radial decomposition for stochastic boundary data in this model. This paper can be used for simplifying the algorithmic use of the spheric-radial decomposition.
In this paper, we analyze the structure of the set of feasible loads in stationary gas networks.
Gas transport through a pipeline network has been the topic of many articles in the last years. These models are often based on the Euler equations (see [4]) or simplifications like the isothermal Euler equations (see [1,2,9,10,11,12,13]). An overview about existing models and the components of a gas network can be found in [5,16]. In [10] the authors show the existence of a unique stationary state, while in [11,14] the model is analyzed for real gas. Optimal control problems with gas networks are considered for example in [3]
It is important to take into account the uncertainty of the boundary data. This leads to optimization problems with probabilistic constraints (see [17]). Especially the model in [9] is quite interesting, because it gives a new access to this topic with respect to uncertain boundary data. This model has also been studied and extended in [12]. The aim in these works is stochastic optimization resp. to answer the question, how large is the probability for a random Gaussian distributed load vector to be feasible. The main tool in these works to compute this probability is the so-called spheric-radial decomposition (see e.g. [21,22,6,9,8]).
Theorem 1.1. (spheric-radial decomposition, see [9], Theorem 2) Let
P(ξ∈M)=∫Sn−1μχ{r≥0|rLv∈M}dμη(v), | (1) |
where
For optimization techniques, it is always an advantage to know, that the admissible set has a special structure like convexity. For example using the integral in Theorem 1.1 for a set
The fact that star-shapedness is an important property in this context, was already mentioned in Assumption 2.1 (ⅱ) in [19]. There star-shapedness is required to define a radial function which maps a ray to the intersection of the ray and a given set. Then if the rays and the given set intersect transversally (cf. Assumption 2.2 (ⅲ), [19]), the Implicit Function Theorem can be applied to this radial function and gradients of probability functions can be computed. Convexity of a given set implies this transversal intersection, if the mean of the Gaussian distribution is in the interior of the given set. Thus convexity can be important for computing gradients of probability functions, e.g. in [23] convexity is a general assumption.
In a nutshell, we analyze the set of feasible loads in the mathematical model of gas transport, depending on the topology of the graph and on the pressure bounds at the nodes. In [20], the authors also analyze the structure of feasible sets in the context of gas transport, though in a different way. The main difference of their model is, that they use a mixed-integer flow model for networks with compressor stations. However, our results about convexity and star-shapedness for networks with compressor stations are not stated in [20]. We will illustrate the difference of the results in the appropriate sections. In Section 2, we shortly introduce the mathematical model. In Section 3, we give a result about convexity of a graph with and without compressor edges and in Section 4 we give some results about when the set of feasible loads is star-shaped to some point.
We will use the model introduced in [12], which is an extension of the model in [9]. The difference between these models is, that the model in [9] does not support compressor stations, which are an important element in gas transport. Compressor stations counteract the pressure drop along the pipes caused by friction. The model in [12] supports these elements.
Consider a connected, directed graph
Definition 2.1. Consider the connected, directed graph
σ(v,e):={−1ife∈E0(v)andf(e)=v1ife∈E0(v)andh(e)=v0ife∉E0(v) |
is called the incidence matrix of the graph
The results of this paper mainly depend on the topology of the graph, so we introduce different structures for a graph:
Definition 2.2. Consider the connected, directed graph
Note, that linear graphs are also tree-structured. Different types of graphs are shown in Figure 1.
For this paper we consider a connected, directed, tree-structured graph
As notation, we state
Corollary 1. For a connected, directed, tree-structured graph, the incidence matrix
Because of the triangular structure of the incidence matrix in tree-structured graphs, the matrices are invertible. Using the Gaussian elimination, the following can be shown:
Corollary 2. The inverse of the incidence matrix
Let
A+q=b+resp.Aq=b. | (2) |
The pressure drop in the flux edges
p2f(e)−p2h(e)=ϕe|qe|qe, | (3) |
and for the compressor edges
(ph(e)pf(e))2=ue, | (4) |
where
Now, we are interested in a solution of this model. The question that we consider is: For a given load vector
M={b+∈Rn+1|1Tn+1b+=0 and ∃(p+,q)∈Rn+1×Rm:p+∈[p+,min,p+,max] and (2),(3),(4)are fulfilled}. | (5) |
Thus, a vector
In general, it is not easy to see, when this set is nonempty. For a given load vector, one has to find a pressure and a flow vector to show this set is nonempty. Obviously a solution of this model might not be unique, there may exist more solutions. For characterizing the set of feasible loads for a tree-structured graph with no compressor edges, we define a function
g:Rn→Rn,g:b↦(AT)−1Φ|A−1b|(A−1b) | (6) |
The matrix
Theorem 2.3. (see [9], Corollary 1) If the network is a tree with a single entry as its root, then the set of feasible nominations is given by
M={(−1Tb,b)∈R−×Rn+|(pmin0)2≤mink=1,⋯,n[(pmaxk)2+gk(b)](pmax0)2≥maxk=1,⋯,n[(pmink)2+gk(b)]maxk=1,⋯,n[(pmink)2+gk(b)]≤mink=1,⋯,n[(pmaxk)2+gk(b)]|} | (7) |
A complete proof of Theorem 2.3 can be found in [9]. For tree-structured networks with compressor stations, the authors of [12] state a similar characterization. The idea here is to separate a graph with
In addition, we define
Theorem 2.4. (see [12], Theorem 5) For given pressure bounds
A vector
(pmini,0)2≤mink=1,⋯,ni[(pmaxi,k)2+gi,k(˜bi)], | (8) |
(pmaxi,0)2≥maxk=1,⋯,ni[(pmini,k)2+gi,k(˜bi)], | (9) |
maxk=1,⋯,ni[(pmini,k)2+gi,k(˜bi)]≤mink=1,⋯,ni[(pmaxi,k)2+gi,k(˜bi)]. | (10) |
For all
1Πk∗,i(pmini,0)2+Σk∗,i(˜b)≤1Πk∗,j(pmaxj,0)2+Σk∗,j(˜b), | (11) |
1Πk∗,i(pmaxi,0)2+Σk∗,i(˜b)≥1Πk∗,j(pminj,0)2+Σk∗,j(˜b), | (12) |
1Πk∗,i(pmini,0)2+Σk∗,i(˜b)≤1Πk∗,jmink=1,⋯,nj[(pmaxj,k)2+gj,k(˜bj)]+Σk∗,j(˜b), | (13) |
1Πk∗,i(pmaxi,0)2+Σk∗,j(˜b)≥1Πk∗,jmaxk=1,⋯,nj[(pminj,k)2+gj,k(˜bj)]+Σk∗,j(˜b), | (14) |
1Πk∗,imaxk=1,⋯,ni[(pmini,k)2+gi,k(˜bi)]+Σk∗,i(˜b)≤1Πk∗,j(pmaxj,0)2+Σk∗,j(˜b), | (15) |
1Πk∗,imink=1,⋯,ni[(pmaxi,k)2+gi,k(˜bi)]+Σk∗,i(˜b)≥1Πk∗,j(pminj,0)2+Σk∗,j(˜b), | (16) |
1Πk∗,imaxk=1,⋯,ni[(pmini,k)2+gi,k(˜bi)]+Σk∗,i(˜b)≤1Πk∗,jmink=1,⋯,nj[(pmaxj,k)2+gj,k(˜bj)]+Σk∗,j(˜b), | (17) |
1Πk∗,imink=1,⋯,ni[(pmaxi,k)2+gi,k(˜bi)]+Σk∗,i(˜b)≥1Πk∗,jmaxk=1,⋯,nj[(pminj,k)2+gj,k(˜bj)]+Σk∗,j(˜b). | (18) |
The values
Σk∗,i(˜b):=n∗−2∑k=11m∗−k∏ℓ=1u(k∗,i),ℓg(k∗,i),n∗−k,f(eu(k∗,i),n∗−k)(˜b(k∗,i),n∗−k)+g(k∗,i),1,f(eu(k∗,i),1)(˜b(k∗,i),1) | (19) |
and
Πk∗,i:=m∗∏k=1u(k∗,i),k. | (20) |
A complete proof of Theorem 2.4 can be found in [12]. The sum defined in (19) as a combination of pressure loss functions and controls, states the change in pressure along a path between subgraphs, e.g.
Here, we will show that in special cases, the feasible set
Lemma 3.1. Let
gk(λb+(1−λ)β)=λ2gk(b)+(1−λ)2gk(β)+2(λ−λ2)Σk(b,β), |
with
Σk(b,β)=k∑i=1ϕi(n∑j=ibj)(n∑j=iβj). | (21) |
Proof. From Corollary 2 we know, that
g(λb+(1−λ)β)=(AT)−1Φ(A−1(λb+(1−λ)β))2. |
The square has to be understood component-by-component. We have
(A−1(λb+(1−λ)β))2= [(n∑i=1(λbi+(1−λ)βi))⋮(n∑i=n(λbi+(1−λ)βi))]2== λ2[(n∑i=1bi)⋮(n∑i=nbi)]2+(1−λ)2[(n∑i=1βi)⋮(n∑i=1βi)]2+2λ(1−λ)[(n∑i=1bi)(n∑i=1βi)⋮(n∑i=nbi)(n∑i=1βi)]. |
We fix a
((A−1)TΦ)ij={ϕjif i≥j0else, |
we get
=λ2k∑i=1ϕi(n∑j=ibj)2+(1−λ)2k∑i=1ϕi(n∑j=iβj)2+2λ(1−λ)k∑i=1ϕi(n∑j=ibj)(n∑j=iβj), |
which is equivalent to
gk(λb+(1−λ)β)=λ2gk(b)+(1−λ)2gk(β)+2λ(1−λ)k∑i=1ϕi(n∑j=ibj)(n∑j=iβj), |
by using the definition of the function
We call the term
Lemma 3.2. With the setting of Lemma 3.1, at least one of the following estimates hold:
(i)Σk(b,β)≤gk(b),(ii)Σk(b,β)≤gk(β). |
Proof. For this proof, we use a classical contradiction argument. We suppose that
Σk(b,β)>gk(b)andΣk(b,β)>gk(β), |
and thus it holds
2Σk(b,β)>gk(b)+gk(β). | (22) |
Since
2Σk(b,β)=2k∑i=1ϕi(n∑j=ibj)(n∑j=iβj)≤k∑i=1ϕi(n∑j=ibj)2+k∑i=1ϕi(n∑j=iβj)2. |
Now the terms on the right are equal to
2Σk(b,β)≤gk(b)+gk(β), | (23) |
which is a contradiction to (22). Thus Lemma 3.2 is proven.
In the last auxiliary lemma, we prove an estimate for a difference of rest terms.
Lemma 3.3. With the setting of Lemma 3.1 and numbers
(i)Σk(b,β)−Σℓ(b,β)≥gk(b)−gℓ(b),(ii)Σk(b,β)−Σℓ(b,β)≥gk(β)−gℓ(β). |
Proof. We use again an contradiction argument to prove this statement. Suppose, that
Σk(b,β)−Σℓ(b,β)<gk(b)−gℓ(b)andΣk(b,β)−Σℓ(b,β)<gk(β)−gℓ(β), |
then we have
2(Σk(b,β)−Σℓ(b,β))<gk(b)−gℓ(b)+gk(β)−gℓ(β). | (24) |
For the left term, we have
2(Σk(b,β)−Σℓ(b,β))==2(k∑i=1ϕi(n∑j=ibj)(n∑j=iβj)−ℓ∑i=1ϕi(n∑j=ibj)(n∑j=iβj)), |
and because
2(Σk(b,β)−Σℓ(b,β))=−2ℓ∑i=k+1ϕi(n∑j=ibj)(n∑j=iβj). |
For the right term in (24) we have
gk(b)−gℓ(b)+gk(β)−gℓ(β)= k∑i=1ϕi(n∑j=1bj)2−ℓ∑i=1ϕi(n∑j=1bj)2+ k∑i=1ϕi(n∑j=1βj)2−ℓ∑i=1ϕi(n∑j=1βj)2, |
and again because
gk(b)−gℓ(b)+gk(β)−gℓ(β)=−(ℓ∑i=k+1ϕi(n∑j=ibj)2+ℓ∑i=k+1(n∑j=iβj)2). |
Now, from a binomial formula, we know that
2(Σk(b,β)−Σℓ(b,β))≥gk(b)−gℓ(b)+gk(β)−gℓ(β), |
which is a contradiction to (24). Thus the lemma is proven.
With these auxiliary lemmas we can state the following convexity Theorem. The result is similar to Lemma 3.3 in [20]. But for that result, the authors only consider the conservation of mass (2), not the conservation of momentum (3). For the model considering both, conservation of mass and momentum, they only state, that the set of feasible loads in general is non-convex (see Lemma 4.1 in [20]).
Theorem 3.4. Let pressure bounds
Proof. First note, that if the feasible set is empty or contains one element, it is convex. Otherwise, it contains at least two elements. In particular we will use the representation of the feasible set
1T(λb++(1−λ)β+)=λ1Tb++(1−λ)1Tβ+=0. | (25) |
Now we have to show, that the following inequalities (see (7)) for convex combinations of loads for
0≤(pmaxk)2−(pmin0)2+gk(λb+(1−λ)β),0≤(pmax0)2−(pmink)2−gk(λb+(1−λ)β),0≤(pmaxk)2−(pminℓ)2+gk(λb+(1−λ)β)−gℓ(λb+(1−λ)β). | (26) |
To show the first inequality in (26) we define
t1,k(b,β)=c1,k+λ2gk(b)+(1−λ)2gk(β)+2(λ−λ2)Σk(b,β). |
Because
t1,k(b,β)≥2(λ−λ2)c1,k+2(λ−λ2)Σk(b,β). |
Now we need an estimate of the form
t1,k(b,β)≥0, |
which implies the first inequality in (26).
We define
t2,k(b,β)=c2,k+λ2gk(b)+(1−λ)2gk(β)+2(λ−λ2)Σk(b,β). |
Next because
t2,k(b,β)≥2(λ−λ2)c2,k−2(λ−λ2)Σk(b,β). |
Now we use the result of Lemma 3.2 to get either
t2,k(b,β)≥2(λ−λ2)c2,k−2(λ−λ2)gk(b),ort2,k(b,β)≥2(λ−λ2)c2,k−2(λ−λ2)gk(β). |
Using again the second inequality in Theorem 2.3, both cases lead to
t2,k(b,β)≥2(λ−λ2)c2,k−2(λ−λ2)c2,k=0, |
which is obviously non-negative and thus, we have shown the second inequality in (26). We define
t3,k,ℓ(b,β)=c3,k,ℓ+λ2(gk(b)−gℓ(b))+(1−λ)2(gk(β)−gℓ(β))+2(λ−λ2)(Σk(b,β)−Σℓ(b,β)). |
Again because
t3,k,ℓ(b,β)≥2(λ−λ2)c3,k,ℓ+2(λ−λ2)(Σk(b,β)−Σℓ(b,β)). |
If
t3,k,ℓ(b,β)≥2(λ−λ2)(c3,k,ℓ+(gk(b)−gℓ(b))),ort3,k,ℓ(b,β)≥2(λ−λ2)(c3,k,ℓ+(gk(β)−gℓ(β))). |
In both cases, we can use the third inequality in Theorem 2.3. It follows
t3,k,ℓ(b,β)≥2(λ−λ2)(c3,k,ℓ−c3,k,ℓ), |
which is non-negative. Finally we have shown
Remark 1. The assumption
We illustrate the results of Theorem 3.4 in the following examples.
Example 1: Consider the linear graph shown in Figure 4.
We consider three different pressure bounds to show the results of Theorem 3.4. In the first case, we consider
Ma{b∈R2≥0|b1≤−b2+√3−b22},Mb{b∈R2≥0|b1≤−b2+√8−b22b2≤√3},Mc{b∈R2≥0|b1≥−b2+√2.25−b22b1≤−b2+√6.75b1≤−b2+√8−b22}. |
The feasible sets are shown in Figure 5.
One can see, that the feasible sets of case
Example 2: Consider the linear graph shown in Figure 6.
We consider two different pressure bounds to show the results of Theorem 3.4. In the first case, we consider
Ma{b∈R3≥0|(b1+b2+b3)2+(b2+b3)2+b23≥4},Ma{b∈R3≥0|(b1+b2+b3)2+(b2+b3)2+b23≥4(b2+b3)2+b23≥1.75(b1+b2+b3)2≤5(b2+b3)2+b23≤5.25(b1+b2+b3)2+(b2+b3)2+b23≤8}. |
The feasible sets are shown in Figure 7 and Figure 8.
In Figure 7 one can see, that the feasible set
Before we give a statement about convexity in a graph with compressor edges, we shortly explain the idea how to treat graphs with compressor edges. The main idea is to remove the compressor edges from the graph, but still keep the properties of the compressors. That means we separate a graph with
Lemma 3.5. For vectors
Σk∗,i(λb+(1−λ)β)=λ2Σk∗,i(b)+(1−λ)2Σk∗,i(β)+2(λ−λ2)Σk∗,i(b,β). |
The sum
Σk∗,i(b,β)=(i−1∑k=k∗+11k−1∏ℓ=k∗uℓnk∑j=1ϕk,j(nk∑α=jbk,α)(nk∑α=jβk,α))+nk∗∑j=1ϕk∗,j(nk∗∑α=jbk∗,α)(nk∗∑α=jβk∗,α) |
Note, that for
Proof of Lemma 3.5. For
Σk∗,i(λb+(1−λ)β)=i−1∑k=k∗+11k−1∏ℓ=k∗uℓgk,nk(λbk+(1−λ)βk)+gk∗,nk∗(λbk∗+(1−λ)βk∗). |
Then from Lemma 3.1 it follows
Σk∗,i(λb+(1−λ)β)=i−1∑k=k∗+11k−1∏ℓ=k∗uℓ(λ2gk,nk(bk)+(1−λ)2gk,nk(βk) |
+ 2(λ−λ2)nk∑j=1ϕk,j(nk∑α=jbk,α)(nk∑α=jβk,α))+(λ2gk∗,nk∗(bk∗)+(1−λ)2gk∗,nk∗(βk∗)+ 2(λ−λ2)nk∗∑j=1ϕk∗,j(nk∗∑α=jbk∗,α)(nk∗∑α=1βk∗,α)), |
which can be written as
Σk∗,i(λb+(1−λ)β)=λ2((i−1∑k=k∗+11k−1∏ℓ=k∗uℓgk,nk(bk))+gk∗,nk∗(bk∗))+ (1−λ)2((i−1∑k=k∗+11k−1∏ℓ=k∗uℓgk,nk(βk))+gk∗,nk∗(βk∗))+ 2(λ−λ2)((i−1∑k=k∗+11k−1∏ℓ=k∗uℓnk∑j=1ϕk,j(nk∑α=jbk,α)(nk∑α=jβk,α))+nk∗∑j=1ϕk∗,j(nk∗∑α=jbk∗,α)(nk∗∑α=jβk∗,α)). |
With the definition of the sum in (19) it follows
Σk∗,i(λb+(1−λ)β)=λ2Σk∗,i(b)+(1−λ)2Σk∗,i+2(λ−λ2)Σk∗,i(b,β), |
and thus the lemma is proven.
For the next convexity theorem, we need a second auxiliary lemma.
Lemma 3.6. With the setting of Lemma 3.5, at least one of the following estimates hold:
Σk∗,i(b,β)≤Σk∗,i(b),orΣk∗,i(b,β)≤Σk∗,i(β). |
Proof. The proof is similar to the proof of Lemma 3.2.
With these results, we can formulate a Theorem about convexity in general linear graphs.
Theorem 3.7. Consider a linear graph
(pmaxi,0)2≥ui−1(pmaxi−1,0)2 | (27) |
hold for
Proof. The proof is similar to the proof of Theorem 3.4. First mention, that if the set of feasible loads is empty or contains only one element, it is convex. Otherwise, let
c4,i,j:=1Πk∗,j(pmaxj,0)2−1Πk∗,i(pmini,0)2, |
with
1Πk∗,j(pmaxj,0)2≥1Πk∗,juj−1(pmaxj−1,0)2≥⋯≥1Πk∗,j(j−1∏k=k∗uk)(pmaxk∗,0)2. |
With the definition of
1Πk∗,j(pmaxj,0)2≥(pmaxk∗,0)2. |
Thus we have
t4,i,j(b,β)=c4,i,j+Σk∗,j(λb+(1−λ)β). |
Then we can use Lemma 3.5 to get
t4,i,j(b,β)=c4,i,j+λ2Σk∗,j(b)+(1−λ)2Σk∗,j(β)+2(λ−λ2)Σk∗,j(b,β). |
Now we use inequality (11) itself for the feasible vectors
t4,i,j(b,β)≥c4,i,j+λ2(−c4,i,j)+(1−λ)2(−c4,i,j)+2(λ−λ2)Σk∗,j(b,β), |
and from this it follows
t4,i,j(b,β)≥2(λ−λ2)c4,i,j+2(λ−λ2)Σk∗,j(b,β). |
Because
c5,i,j:=1Πk∗,i(pmaxi,0)2−1Πk∗,j(pmink∗,j)2, |
and
t5,i,j(b,β):=c5,i,j−Σk∗,j(λb+(1−λ)β). |
Because
t5,i,j(b,β)=c5,i,j−(λ2Σk∗,j(b)+(1−λ)2Σk∗,j(β)+2(λ−λ2)Σk∗,j(b,β)). |
Then, we use the inequality (12) itself for feasible
t5,i,j(b,β)≥c5,i,j+λ2(−c5,i,j)+(1−λ)2(−c5,i,j)−2(λ−λ2)Σk∗,j(b,β), |
from which it follows
t5,i,j(b,β)≥2(λ−λ2)c5,i,j−2(λ−λ2)Σk∗,j(b,β). |
Now we use Lemma 3.6. It follows either
t5,i,j(b,β)≥2(λ−λ2)c5,i,j−2(λ−λ2)Σk∗,j(b),ort5,i,j(b,β)≥2(λ−λ2)c5,i,j−2(λ−λ2)Σk∗,j(β). |
In both cases, we can use again inequality (12) to get
t5,i,j(b,β)≥2(λ−λ2)c5,i,j−2(λ−λ2)c5,i,j, |
which is obviously non-negative. Thus inequality (12) holds for
Remark 2. The extra condition
The following example illustrates the results.
Example 3. : Consider the minimal graph with a compressor edge shown in Figure 9.
We consider two different controls for pressure bounds
Ma{b∈RR≥03|(b1+b2+b3)2≤8b23≤3(b1+b2+b3)2+12b23≤8.75},Mb{b∈R3≥0|(b1+b2+b3)2≤8b23≤3(b1+b2+b3)2+14b23≤8.75(b1+b2+b3)2+14b23≥0.4375}. |
The feasible sets are shown in Figure 10 and Figure 11.
In both cases, one can see that the feasible sets are convex. Even if in case
In this section, we will show that for tree-structured graphs, the feasible set is star-shaped to some special points. Like it is mentioned in Section 1, computing the intervals, in which a line through a fixed point intersects the set of feasible loads, is much easier if one knows that the set is star-shaped to this point. First we show the following auxiliary lemma.
Lemma 4.1. Let
g(λb)=λ2g(b). | (28) |
Proof. Consider
g(λb)=(AT)−1Φ|A−1λb|(A−1λb)=(AT)−1Φ|λ||A−1b|λ(A−1b)=λ2(AT)−1Φ||A−1b|(A−1b)=λ2g(b), |
since
Now we formulate a theorem about when the set of feasible loads is star-shaped. This result is equal to {Lemma 4.2} in [20]. However, we state this theorem here because we prove it differently and we will use the proof for a similar result for networks with compressor stations later. As mentioned in Section 1, the main difference between the models is the modeling of the compressor stations.
Theorem 4.2. Let pressure bounds
Proof. Let
c1,i:=(pmaxi)2−(pmin0)2, |
and
t1,i(b):=c1,i+gi(λb). |
Then with Lemma 4.1 we have
t1,i(b)=c1,i+λ2gi(b). |
Because
t1,i(b)≥c1,i−λ2c1,i. |
Because
c2,i:=(pmax0)2−(pmini)2,andt2,i(b):=c2,i−gi(λb). |
Then due to Lemma 4.1 we have
t1,i(b)=c1,i−λ2gi(b). |
We use the second inequality of Theorem 2.3 and get
t1,i(b)≥c1,i−λ2c1,i, |
and it follows
c3,i,j:=(pmaxi)2−(pminj)2,andt3,i,j(b):=c3,i,j+gi(λb)−gj(λb). |
With Lemma 4.1 it follows
t3,i,j(b)=c3,i,j+λ2(gi(b)−gj(b)). |
We use the third inequality in Theorem 2.3, we have
t3,i,j(b)≥c3,i,j−λ2c3,i,j, |
so it follows
With the next example, we illustrate why we need the statement
Example 4: Consider the minimal tree shown in Figure 12.
We consider the two cases
Mb{b∈R2≥0|b1≤,b2≤√8b21≤3+b22b22≤3+b21},Mb{b∈R2≥0|b2≥1.5b1≤√6.75b2≤√8b21≤1.75+b22b22≤5.25+b21}, |
The feasible sets are shown in Figure 13.
One can see, that in case
Example 5: Consider the minimal tree shown in Figure 14.
We consider the two cases
Ma={b∈R3≥0|b21≤8(b2+b3)2+b23≤8b21≤3+(b2+b3)2(b2+b3)2+b23≤3+b21b23≤3}, |
and Mb={b∈R3≥0|(b2+b3)2+b23≥1.75b21≤8(b2+b3)2+b23≤8b21≤3+(b2+b3)2b21≤1.25+(b2+b3)2+b23(b2+b3)2+b23≤3+b21b23≤3} |
The feasible sets are shown in Figure 15 and Figure 16.
As in Example 4, the feasible set in case
Next we show, that the statement of Theorem 4.2 also holds for tree-structured graphs with compressor edges. This result is not stated in [20].
Theorem 4.3. Let controls
Proof. Consider
Σk∗,i(λb)=λ2Σk∗,i(b). | (29) |
The proof follows the structure of the proof of Theorem 4.2. We set
c4,i,j:=1Πk∗,j(pmaxj,0)2−1Πk∗,i(pmini,0)2,t4,i,j(b):=c4,i,j+Σk∗,j(λb)−Σk∗,i(λb), |
and
c5,i,j:=1Πk∗,i(pmaxi,0)2−1Πk∗,j(pminj,0)2,t5,i,j(b):=c5,i,j+Σk∗,i(λb)−Σk∗,j(λb). |
Then with (29) we have
t4,i,j(b)=c4,i,j+λ2(Σk∗,j(b)−Σk∗,i(b)), |
and
t5,i,j(b)=c5,i,j+λ2(Σk∗,i(b)−Σk∗,j(b)). |
Because
t4,i,j(b)≥c4i,j+λ2(−c4i,j), |
and
t5,i,j(b)≥c5i,j+λ2(−c5i,j). |
From Theorem 3.7 we know that
The last statement in this section is motivated by case b) in Example 4 (see Figure 13). Because we want to know, when the intersection of a line and the set of feasible loads is convex, we formulate the following lemma.
Lemma 4.4. Let
Remark 3. The statement of Lemma 4.4 is a generalized star-shapedness property with respect to the point
Proof of Lemma 4.4. First, if the set
Part I: Proof for trees without compressor edges: We first prove this lemma for trees without compressor edges. Consider
(pmin0)2−(pmaxk)2≤gk(β1), |
which is equal to
(pmin0)2−(pmaxk)2≤gk(λ1b). |
From Lemma 4.1, it follows
(pmin0)2−(pmaxk)2≤λ21gk(b), |
and this also holds for every
(pmax0)2−(pmin0)2≥gk(β2). |
This is equal to
(pmax0)2−(pmin0)2≥gk(λ2b), |
and due to Lemma 4.1 it follows
(pmax0)2−(pmin0)2≥λ22gk(b). |
This also holds for
(pmink)2−(pmaxℓ)2≤gk(β1)−gℓ(β1),resp.(pmink)2−(pmaxℓ)2≤gk(β2)−gℓ(β2). |
The term on the right in both cases has the same sign, because if
Part II: Proof for general trees: Now we prove this lemma for general trees with compressor edges. This part of the proof follows the structure of Part I. Consider
1Πk∗,i(pmini,0)2−1Πk∗,j(pmaxj,0)2≤Σk∗j(βk)−Σk∗,i(βk)(k=1,2). |
We follow the argumentation of the third inequality in Part I of the proof. The term on the right is either negative or non-negative. If it is non-negative, we follow the argumentation of the first inequality in Part I, if it is negative, we follow the argumentation of the second inequality in Part I. We can use the same arguments to show the inequalities (12) - (18) for
The statement of Lemma 4.4 is mentioned and illustrated in Example 4 and Example 5. This is listed here, because it also belongs to the problems mentioned in Section 1.
In this paper we have shown that the structure of the set of feasible loads in the context of stationary gas networks mainly depends on the topology of the network graphs. In the easiest case, if the network graph is linear, convexity of the set of feasible loads can be shown with very weak assumptions. If the network graph is more complex, but still does not contain circles, we have shown, that under weak assumptions, the feasible set is always star-shaped with respect to the point
In [9], also the case of a network, which is a cycle, is considered. Then, the representation of the set of feasible loads shown in Theorem 2.3 is completed with an equality. For this framework, one can also show, that the set of feasible loads is star-shaped to the point
As mentioned in Section 1, knowing the structure of the set of feasible loads helps to analyze this model for random load vectors as it is done in [9] and [12]. There, the authors use the spheric-radial decomposition to handle the probabilistic load vectors. This leads to optimization problems with probabilistic constrains or chance constraints (see [17]). The main part of using the spheric-radial decomposition (see Theorem 1.1) is to compute the integral in (1). If one knows e.g. that the feasible set is convex, this integral is a lot easier to compute and thus, this simplifies the optimization done in [12].
The authors declare that there is no conflict of interest regarding the publication of this paper.
This work was supported by DFG in the framework of the Collaborative Research Centre CRC/Transregio 154, Mathematical Modelling, Simulation and Optimization Using the Example of Gas Networks, project C03.
[1] |
Coupling conditions for gas networks governed by the isothermal Euler equations. Netw. Heterog. Media (2006) 1: 295-314. ![]() |
[2] |
Gas flow in pipeline networks. Netw. Heterog. Media (2006) 1: 41-56. ![]() |
[3] | Simulation and optimization models of steady-state gas transmission networks. Energy Procedia (2015) 64: 130-139. |
[4] |
Euler systems for compressible fluids at a junction. J. Hyperbolic Differ. Equ. (2008) 5: 547-568. ![]() |
[5] | P. Domschke, B. Hiller, J. Lang and C. Tischendorf, Modellierung von Gasnetzwerken: Eine Übersicht, (2019), Preprint on Webpage http://tubiblio.ulb.tu-darmstadt.de/106763/. |
[6] |
Properties of chance constraints in infinite dimensions with an application to PDE constrained optimization. Set-Valued Var. Anal. (2018) 26: 821-841. ![]() |
[7] |
A. Genz and F. Bretz, Computation of Multivariate Normal and t Probabilities, Lecture Notes in Statistics, 195. Springer, Dordrecht, 2009. doi: 10.1007/978-3-642-01689-9
![]() |
[8] |
A joint model of probabilistic/robust constraints for gas transport management in stationary networks. Comput. Manag. Sci. (2017) 14: 443-460. ![]() |
[9] |
On the quantification of nomination feasibility in stationary gas networks with random load. Math. Methods Oper. Res. (2016) 84: 427-457. ![]() |
[10] |
Stationary states in gas networks. Netw. Heterog. Media (2015) 10: 295-320. ![]() |
[11] |
Networks of pipelines for gas with nonconstant compressibility factor: Stationary states. J. Comput. Appl. Math. (2018) 37: 1066-1097. ![]() |
[12] |
M. Gugat and M. Schuster, Stationary gas networks with compressor control and random loads: Optimization with probabilistic constraints, Math. Probl. Eng., 2018 (2018), Art. ID 7984079, 17 pp. doi: 10.1155/2018/7984079
![]() |
[13] |
The isothermal Euler equations for ideal gas with source term: Product solutions, flow reversal and no blow up. J. Math. Anal. Appl. (2017) 454: 439-452. ![]() |
[14] |
Transient flow in gas networks: Traveling waves. Int. J. Appl. Math. Comput. Sci. (2018) 28: 341-348. ![]() |
[15] |
H. Heitsch, On probabilistic capacity maximization in a stationary gas network, Optimization, 69 (2020), 575–604, Preprint on Webpage http://wias-berlin.de/publications/wias-publ/run.jsp?template=abstract&type=Preprint&year=2018&number=2540. doi: 10.1080/02331934.2019.1625353
![]() |
[16] |
T. Koch, B. Hiller, M. E. Pfetsch and L. Schewe, Evaluating Gas Network Capacities, MOS-SIAM Series on Optimization, 21. Society for Industrial and Applied Mathematics (SIAM), Philadelphia, PA, 2015. doi: 978-1-611973-68-6
![]() |
[17] |
A. Prékopa, Stochastic Programming, Mathematics and its Applications, 324. Kluwer Academic Publishers Group, Dordrecht, 1995. doi: 10.1007/978-94-017-3087-7
![]() |
[18] |
Computational optimization of gas compressor stations: MINLP models versus continuous reformulations. Math. Methods Oper. Res. (2016) 83: 409-444. ![]() |
[19] |
Extensions of stochastic optimization results to problems with system failure probability functions. J. Optim. Theory Appl. (2007) 133: 1-18. ![]() |
[20] |
L. Schewe, M. Schmidt and J. Thürauf, Structural properties of feasible bookings in the European entry-exit gas market system, 4OR, (2019). doi: 10.1007/s10288-019-00411-3
![]() |
[21] |
(Sub-)Differentiability of probabilistic functions with elliptical distributions. Set-Valued Var. Anal. (2018) 26: 887-910. ![]() |
[22] |
Gradient formulae for nonlinear probabilistic constraints with Gaussian and Gaussian-like distribution. SIAM J. Optim. (2014) 24: 1864-1889. ![]() |
[23] | D. Wintergerst, Application of chance constrained optimization to gas networks, (2019), Preprint on Webpage https://opus4.kobv.de/opus4-trr154/frontdoor/index/index/start/5/rows/10/sortfield/score/sortorder/desc/searchtype/simple/query/wintergerst/docId/158. |
1. | Michael Schuster, Elisa Strauch, Martin Gugat, Jens Lang, Probabilistic constrained optimization on flow networks, 2022, 23, 1389-4420, 1, 10.1007/s11081-021-09619-x | |
2. | Martin Gugat, Michael Herty, 2022, 23, 9780323850599, 59, 10.1016/bs.hna.2021.12.002 | |
3. | María Concepción López-Díaz, Miguel López-Díaz, On the Axial Symmetry of 2D Star-Shaped Sets, 2025, 67, 0924-9907, 1, 10.1007/s10851-024-01222-w | |
4. | Martin Gugat, Michael Schuster, Jan Sokołowski, The location problem for compressor stations in pipeline networks, 2024, 12, 2325-3444, 507, 10.2140/memocs.2024.12.507 | |
5. | Andrea Corli, Massimiliano D. Rosini, Ulrich Razafison, 2024, Mathematical Modeling of Chattering and the Optimal Design of a Valve*, 979-8-3503-1633-9, 76, 10.1109/CDC56724.2024.10886245 |