Loading [MathJax]/jax/output/SVG/jax.js

Convexity and starshapedness of feasible sets in stationary flow networks

  • Received: 01 January 2019 Revised: 01 October 2019 Published: 01 June 2020
  • Primary: 90C15; Secondary: 93E20

  • 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

    Related Papers:

    [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 ξN(0,R) be the n-dimensional standard Gaussian distribution with zero mean and positive definite correlation matrix R. Then, for any Borel measurable subset MRn it holds that

    P(ξM)=Sn1μχ{r0|rLvM}dμη(v), (1)

    where Sn1 is the (n1)-dimensional sphere in Rn, μη is the uniform distribution on Sn1, μχ denotes the χ-distribution with n degrees of freedom and L is such that R=LLT (e.g., Cholesky decomposition).

    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 M, a sampled point vSn1 and a square matrix LRn (which is a Cholesky decomposition of the covariance matrix of the distribution), one has to compute the one-dimensional set {r0 | rLvM}. This set can be represented as a union of disjoint intervals, but for big graphs, the number of disjoint intervals can be very large. So the numerical computation of this union can be very time-demanding. The idea of this paper is, that knowledge about the structure of set M, which is the set of feasible loads in our model, allows to reduce the time of computation enormously. E.g. if one know that M is convex or star-shaped with respect to some point, the sets {r0 | rLvM} are just convex intervals. This implies that the algorithm for computing these unions of disjoint intervals can stop as soon as it finds one interval.

    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 G=(V+,E) which represents a pipeline gas transport network. We set |V+|=n+1 and |E|=m (m,nN). We introduce the following notation for graphs:

    Definition 2.1. Consider the connected, directed graph G=(V+,E):

    (i) h(e) denotes the head node of an edge e and f(e) denotes the foot node of an edge e for all eE

    (ii) E0(v):={eE|h(e)=v or f(e)=v} denotes all edges which are connected to node vV+

    (iii) The matrix A+i,jRn+1×m, A+i,j=σ(vi,ej) with

    σ(v,e):={1ifeE0(v)andf(e)=v1ifeE0(v)andh(e)=v0ifeE0(v)

    is called the incidence matrix of the graph G.

    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 G=(V+,E):

    (i) & The graph G is called linear, if E(v0)=1 and |E(v)|2 for all vV+

    (ii) & The graph G is called tree-structured, if there exists no edge eE with h(e)=v0 and for all nodes vV+ there is at most one edge eE with h(e)=v.

    Note, that linear graphs are also tree-structured. Different types of graphs are shown in Figure 1.

    Figure 1.  Example of differently structured graphs.

    For this paper we consider a connected, directed, tree-structured graph G=(V+,E) with |V+|=n+1 nodes and |E|=m=n edges. We assume, that the root of the tree is the only influx node (gas enters the network) and all other nodes are efflux nodes (gas leaves the network). An edge can either be a flux edge, so the pressure decreases along the edge, or a compressor edge, so the pressure increases along the edge. We define EF as the set of all flux edges and EC as the set of compressor edges. We have E=EFEC with EFEC=. We determine a numbering for the nodes and edges of the graph. The input node gets the number 0 and all other nodes are numbered using breadth-first search or depth-first search. Every edge eE gets the number max{h(e),f(e)}.

    As notation, we state V=V+{v0} and A as A+ without the first row, which corresponds to the influx node. Then, the incidence matrix A for tree-structured graphs is square. The following fact is easy to see:

    Corollary 1. For a connected, directed, tree-structured graph, the incidence matrix A is an upper triangular matrix with 1 at its diagonal. Moreover, if the graph is linear, the incidence matrix is 1 at its diagonal, 1 at the diagonal above and 0 elsewhere.

    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 A of a connected, directed, tree-structured graph is also upper triangular with values in {0,1}. Moreover, if the graph is linear, the inverse incidence matrix is 1 in the upper triangle and 0 elsewhere.

    Let b+Rn+1 with 1n+1b+=0 denote the balanced load vector and assume bi<0 for the node with gas influx and bi0 for nodes with gas efflux (i{0,,n}), where 1n+1 is the vector of all ones in the dimension n+1. Let qRm denote the flows in the edges and p+Rn+1 denotes the pressures at the nodes. Again we set b resp. p as b+ resp. p+ without the first component, corresponding to the inflow node. For the pressure we consider the constraints p+[p+,min,p+,max]. The conservation of mass for the graph is given by

    A+q=b+resp.Aq=b. (2)

    The pressure drop in the flux edges eEF is given by the so-called Weymouth equation (see e.g. [11])

    p2f(e)p2h(e)=ϕe|qe|qe, (3)

    and for the compressor edges eEC we have

    (ph(e)pf(e))2=ue, (4)

    where ϕe and ue are constants. The compressor stations counteract the pressure drop caused by friction in the pipes. For a more detailed model derivation we refer to [5,12,18].

    Now, we are interested in a solution of this model. The question that we consider is: For a given load vector bRn (in a tree-structured graph), when can we find corresponding vectors of pressure and flow, so that the box constraints for the pressure, the equation for mass conservation, the equation for the pressure drop and the equation for the compressor stations are fulfilled? Therefore we define the set of feasible loads (called feasible set, here M):

    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 bRn of given outflows is feasible if and only if (1Tnb,b)M.

    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:RnRn,g:b(AT)1Φ|A1b|(A1b) (6)

    The matrix ΦRn×n is a diagonal matrix with the values ϕi at its diagonal. The components gk(b) describe the pressure loss from the root to node vk (k=1,,n) with the load vector b. In [9], the authors characterize the set of feasible loads for tree-structured graphs without compressor stations. The idea is as follows: Consider a feasible pressure at a certain node, i.e. the pressure at this node satisfies the box constraints. With the function g, we can follow the change in this pressure along the edges in the graph. We have to guarantee, that the changed pressure also fulfills the box constraints at the other nodes. This is listed in the next theorem.

    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)2mink=1,,n[(pmaxk)2+gk(b)](pmax0)2maxk=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 mc compressor edges in mc+1 subgraphs by removing the compressor edges, but still keep the property of the compressor stations. The subgraphs and the nodes inside every subgraph are numbered by breadth-first search resp. depth-first search. Then the notation is the following: Gi denotes the subgraph with number i, pi,k, bi,k resp. gi,k denote the k-th component of the pressure, load vector resp. pressure loss function of the subgraph Gi and vi,k is the node with number k in Gi. This is shown in Figure 2.

    Figure 2.  Example for illustrating the notation (graph numbered by breadth-first search).

    In addition, we define p(i,j),k, b(i,j),k resp. g(i,j),k as the pressure, load vector resp. pressure loss function, which belongs to the k-th subgraph between Gi and Gj, s.t. p(i,j),1 and b(i,j),1 belong to Gi. Further, p(i,j),k,, b(i,j),k, resp. g(i,j),k, is the -th component of p(i,j),k, b(i,j),k resp. g(i,j),k. Similarly, we set u(i,j),k as k-th control on the path from vi,0 to vj,0. An example of this notation for i=1 and j=5 is shown in Figure 3. We also define ki,j as the largest index of all subgraphs, the paths from the root to vi,0 and vj,0 pass. E.g. we have k2,5=1 and k2,4=2 (cf. Figure 2. Last we define ni,j as the number of subgraphs, the path from vi,0 to vj,0 pass and mi,j as number of controls, the path from vi,0 to vj,0 pass. E.g. it is n1,5=3 and m1,5=2 (cf. Figure 2 and Figure 3). For better readability, we only write k instead of ki,j and n resp. m instead of nk,i resp. mk,i (i,j=1,,n). The notation is also explained in detail in [12] above Theorem 5. Then the idea of guaranteeing feasibility is the same as explained above. In the next theorem, a characterization of the set of feasible loads for tree-structured networks with compressor stations is stated.

    Figure 3.  Example for illustrating the triple and quadruple indices on the path from G1 to G5.

    Theorem 2.4. (see [12], Theorem 5) For given pressure bounds p+,min,p+,maxRn+1 and controls uiR (i=1,,m2) the following equivalence holds:

    A vector b+ with 1Tb+=0 is feasible, i.e. b+M, if and only if the following inequalities hold: For all i=1,,m2+1 holds (feasibility inside the subgraphs)

    (pmini,0)2mink=1,,ni[(pmaxi,k)2+gi,k(˜bi)], (8)
    (pmaxi,0)2maxk=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 i,j=1,,m2+1 with i<j holds (feasibility between the subgraphs)

    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 and Πk,j are defined as

    Σk,i(˜b):=n2k=11mk=1u(k,i),g(k,i),nk,f(eu(k,i),nk)(˜b(k,i),nk)+g(k,i),1,f(eu(k,i),1)(˜b(k,i),1) (19)

    and

    Πk,i:=mk=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. Σ1,4(˜b) gives the chance in pressure between node v1,0 and v2,1 from Figure 2. One can see, that in both theorems (Theorem 2.3 and Theorem 2.4), the decision whether a given load vector is feasible or not only depends on the pressure bounds. So for a given load vector, one has to check if a number of inequalities depending on the load vector and the pressure bounds are fulfilled. This is an enormous simplification to deal with the set of feasible loads. We mention again here, that we distinguish between the load vector b+ (full load vector) and b (load vector without the first component).

    Here, we will show that in special cases, the feasible set M (defined in (5)) is convex. As it is mentioned in Section 1, the computation of the probability for a random Gaussian distributed load vector to be feasible by using the {spheric-radial decomposition}, simplifies a lot if one knows that the set of feasible loads is convex. Also in this case it is possible to use other algorithms (see e.g. [7]). Throughout this section, we assume that our graph is linear, that is tree-structured without any branching. Thus, the numbering using depth-first search equals the numbering using breadth-first search. We first show a few auxiliary lemmas for the pressure loss function (defined for tree-structured graphs without compressor edges in (6)). The first Lemma is about evaluating g() at a convex combination of load vectors.

    Lemma 3.1. Let MRn0 be the set of feasible loads. For b,βM, constants ϕi0 (i=1,,n) and λ(0,1), it holds (for all k=1,,n):

    gk(λb+(1λ)β)=λ2gk(b)+(1λ)2gk(β)+2(λλ2)Σk(b,β),

    with

    Σk(b,β)=ki=1ϕi(nj=ibj)(nj=iβj). (21)

    Proof. From Corollary 2 we know, that A1 contains only non-negative values. The load vectors b and β only corresponds to efflux nodes, so they also contain only non-negative values and thus the pressure loss function can be written as

    g(λb+(1λ)β)=(AT)1Φ(A1(λb+(1λ)β))2.

    The square has to be understood component-by-component. We have

    (A1(λb+(1λ)β))2= [(ni=1(λbi+(1λ)βi))(ni=n(λbi+(1λ)βi))]2== λ2[(ni=1bi)(ni=nbi)]2+(1λ)2[(ni=1βi)(ni=1βi)]2+2λ(1λ)[(ni=1bi)(ni=1βi)(ni=nbi)(ni=1βi)].

    We fix a k{1,,n}. Together with

    ((A1)TΦ)ij={ϕjif ij0else,

    we get gk(λb+(1λ)β)=

    =λ2ki=1ϕi(nj=ibj)2+(1λ)2ki=1ϕi(nj=iβj)2+2λ(1λ)ki=1ϕi(nj=ibj)(nj=iβj),

    which is equivalent to

    gk(λb+(1λ)β)=λ2gk(b)+(1λ)2gk(β)+2λ(1λ)ki=1ϕi(nj=ibj)(nj=iβj),

    by using the definition of the function g (see (6)). With Σk(b,β) defined in (21) the Lemma is proven.

    We call the term Σk(b,β) the remainder term of g evaluated at a convex combination. In the next Lemma, we prove an estimate for this remainder 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 2aba2+b2 for real numbers a and b, with this estimate and the definition of Σk(b,β), it follows

    2Σk(b,β)=2ki=1ϕi(nj=ibj)(nj=iβj)ki=1ϕi(nj=ibj)2+ki=1ϕi(nj=iβj)2.

    Now the terms on the right are equal to gk(b) resp. gk(β) and thus it follows

    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 k,{1,,n} (with k<), at least one of the following estimates hold:

    (i)Σk(b,β)Σ(b,β)gk(b)g(b),(ii)Σk(b,β)Σ(b,β)gk(β)g(β).

    Proof. We use again an contradiction argument to prove this statement. Suppose, that

    Σk(b,β)Σ(b,β)<gk(b)g(b)andΣk(b,β)Σ(b,β)<gk(β)g(β),

    then we have

    2(Σk(b,β)Σ(b,β))<gk(b)g(b)+gk(β)g(β). (24)

    For the left term, we have

    2(Σk(b,β)Σ(b,β))==2(ki=1ϕi(nj=ibj)(nj=iβj)i=1ϕi(nj=ibj)(nj=iβj)),

    and because k< this implies

    2(Σk(b,β)Σ(b,β))=2i=k+1ϕi(nj=ibj)(nj=iβj).

    For the right term in (24) we have

    gk(b)g(b)+gk(β)g(β)= ki=1ϕi(nj=1bj)2i=1ϕi(nj=1bj)2+ ki=1ϕi(nj=1βj)2i=1ϕi(nj=1βj)2,

    and again because k< it follows

    gk(b)g(b)+gk(β)g(β)=(i=k+1ϕi(nj=ibj)2+i=k+1(nj=iβj)2).

    Now, from a binomial formula, we know that 2ab(a2+b2) for real numbers a and b. This leads to the estimate

    2(Σk(b,β)Σ(b,β))gk(b)g(b)+gk(β)g(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 p+,min,p+,maxRn+1 with pmaxipminj (for all i,j=0,,n) be given. Then, for a linear network graph with one single entry and no compressor edges, the set of feasible loads is convex.

    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 M of Theorem 2.3. The formulation (1Tb,b)R×Rn+ is equivalent to 1Tb+=0 for graphs with one single entry. Consider b+,β+M. It holds 1Tb+=0, 1Tβ+=0 and all inequalities in Theorem 2.3 are fulfilled. We have to show, that for a λ(0,1), the load vector λb++(1λ)β+ is also in M. First we have

    1T(λb++(1λ)β+)=λ1Tb++(1λ)1Tβ+=0. (25)

    Now we have to show, that the following inequalities (see (7)) for convex combinations of loads for k,=1,,n hold:

    0(pmaxk)2(pmin0)2+gk(λb+(1λ)β),0(pmax0)2(pmink)2gk(λb+(1λ)β),0(pmaxk)2(pmin)2+gk(λb+(1λ)β)g(λb+(1λ)β). (26)

    To show the first inequality in (26) we define c1,k:=(pmaxk)2(pmin0)2 and t1,k(b,β):=c1,k+gk(λb+(1λ)β) and we want to show t1,k(b,β)0. With Lemma 3.1 it follows

    t1,k(b,β)=c1,k+λ2gk(b)+(1λ)2gk(β)+2(λλ2)Σk(b,β).

    Because b and β are feasible, we can use the first inequality in Theorem 2.3 to get the estimate

    t1,k(b,β)2(λλ2)c1,k+2(λλ2)Σk(b,β).

    Now we need an estimate of the form Σk(b,β)(c1,k), but this is not true in general. Here we use the restriction on the pressure bounds. Because pmaxipminj for all i,j=0,,n, it follows c1,k0 for all k=1,,n. And because Σk(b,β)0, we have

    t1,k(b,β)0,

    which implies the first inequality in (26).

    We define c2,k:=(pmax0)2(pmink)2, t2,k(b,β):=c2,kgk(λb+(1λ)β) and we use again Lemma 3.1 and get

    t2,k(b,β)=c2,k+λ2gk(b)+(1λ)2gk(β)+2(λλ2)Σk(b,β).

    Next because b and β are feasible, we use the second inequality in Theorem 2.3 to get the estimate

    t2,k(b,β)2(λλ2)c2,k2(λλ2)Σk(b,β).

    Now we use the result of Lemma 3.2 to get either

    t2,k(b,β)2(λλ2)c2,k2(λλ2)gk(b),ort2,k(b,β)2(λλ2)c2,k2(λλ2)gk(β).

    Using again the second inequality in Theorem 2.3, both cases lead to

    t2,k(b,β)2(λλ2)c2,k2(λλ2)c2,k=0,

    which is obviously non-negative and thus, we have shown the second inequality in (26). We define c3,k,:=(pmaxk)2(pmin)2 and t3,k,(b,β):=c3,k,+gk(λb+(1λ)β)g(λb+(1λ)β). In the case k= it follows directly t3,k,(b,β)0 (because of c3,k,0 due to assumptions). In the case k, with Lemma 3.1 it follows

    t3,k,(b,β)=c3,k,+λ2(gk(b)g(b))+(1λ)2(gk(β)g(β))+2(λλ2)(Σk(b,β)Σ(b,β)).

    Again because b and β are feasible, we use the third inequality in Theorem 2.3 to get

    t3,k,(b,β)2(λλ2)c3,k,+2(λλ2)(Σk(b,β)Σ(b,β)).

    If <k, it follows Σ(b,β)<Σk(b,β) and thus every term is positive, which implies t3,k,(b,β)0. If >k, we use the statement of Lemma 3.3. With this, it follows either

    t3,k,(b,β)2(λλ2)(c3,k,+(gk(b)g(b))),ort3,k,(b,β)2(λλ2)(c3,k,+(gk(β)g(β))).

    In both cases, we can use the third inequality in Theorem 2.3. It follows

    t3,k,(b,β)2(λλ2)(c3,k,c3,k,),

    which is non-negative. Finally we have shown t3,k,0, so the load vector (λb+(1λ)β) is also feasible and thus the set of feasible loads is convex.

    Remark 1. The assumption pmaxipminj is not a natural restriction, because otherwise, the trivial solution is no element of the set of feasible loads. We will assume this later also to show star-shapedness. It is interesting to see, that this assumption also occurs in the problem of maximizing booked capacities on stationary tree-structured graphs, in verifying constraint qualifications for certain problems (see Theorem 1 in [15]).

    We illustrate the results of Theorem 3.4 in the following examples.

    Example 1: Consider the linear graph shown in Figure 4.

    Figure 4.  Linear graph of example 1.

    We consider three different pressure bounds to show the results of Theorem 3.4. In the first case, we consider (p+,min)a=[2,1,1]T and (p+,max)a=[2,2,2]T with the feasible set Ma, in the second case we consider (p+,min)b=[2,1,1]T and (p+,max)b=[3,2,2]T with feasible set Mb and in the third case, we consider (p+,min)c=[2.5,1.5,1]T and (p+,max)c=[3,2.5,2]T with feasible set Mc. Together with ϕ1=ϕ2=1, we get from Theorem 2.3 the feasible sets

    Ma{bR20|b1b2+3b22},Mb{bR20|b1b2+8b22b23},Mc{bR20|b1b2+2.25b22b1b2+6.75b1b2+8b22}.

    The feasible sets are shown in Figure 5.

    Figure 5.  Feasible sets for different pressure bounds.

    One can see, that the feasible sets of case (a) and (b) are convex and the feasible set of case (c) is not. This fits to the result of Theorem 3.4 because in case (a) and (b), the condition p+,maxip+,minj is fulfilled, but not in case (c).

    Example 2: Consider the linear graph shown in Figure 6.

    Figure 6.  Linear graph of example 2.

    We consider two different pressure bounds to show the results of Theorem 3.4. In the first case, we consider (p+,min)a=[1,1,1,1]T, (p+,max)a=[3,3,3,3]T and in the second case, we consider (p+,min)b=[2.5,2,1.5,1]T, (p+,max)b=[3,2.5,2,1.5]T. With ϕ1=ϕ2=ϕ3=1, Theorem 2.3 implies

    Ma{bR30|(b1+b2+b3)2+(b2+b3)2+b234},Ma{bR30|(b1+b2+b3)2+(b2+b3)2+b234(b2+b3)2+b231.75(b1+b2+b3)25(b2+b3)2+b235.25(b1+b2+b3)2+(b2+b3)2+b238}.

    The feasible sets are shown in Figure 7 and Figure 8.

    Figure 7.  Set Ma for (p+,min)a=[1,1,1,1]T and (p+,max)a=[3,3,3,3]T.
    Figure 8.  Set Mb for (p+,min)b=[2.5,2,1.5,1]T and (p+,max)b=[3,2.5,2,1.5]T.

    In Figure 7 one can see, that the feasible set Ma is convex. In the view from above one can see, that the picture is similar to the picture in the two-dimensional case Figure 5 a). In Figure 8, the feasible set is obviously not convex. This is because the condition pminipmaxj is not fulfilled for every i,j=1,,n. This feasible set is similar to the set in the two dimensional example in Figure 5 (c).

    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 mN compressor edges to m+1 subgraphs, which are not connected, but which interact with each other. We numerate the subgraphs exactly like we numerate the nodes inside a subgraph (with breadth-first search or depth-first search). This is explained in detail in [12], Section 3. We formulate an auxiliary lemma for the sum Σk,i(b) (defined in (19)) first. For an easier notation, we can use the fact, that the graph is linear. The notation in [12] is motivated by paths in tree-structured graphs, but for linear graphs, there exists only one path in the graph. So for the next lemma, we state that gi,j(bi) is the j-th component of the pressure loss function g for the i-th subgraph and bi is the load vector (without the first component) for the i-th subgraph. Further, ni+1 is the number of nodes in the i-th subgraph, numbered from 0 to nk. Analogously, ϕi,j belongs to the edge with number j in the i-th subgraph.

    Lemma 3.5. For vectors b,βRn0, constants ϕkR0 (i=1,,n) and λ(0,1), it holds (for i=1,,n):

    Σk,i(λb+(1λ)β)=λ2Σk,i(b)+(1λ)2Σk,i(β)+2(λλ2)Σk,i(b,β).

    The sum Σk,i() is defined in (19) and Σk,i(b,β) is defined as:

    Σk,i(b,β)=(i1k=k+11k1=kunkj=1ϕk,j(nkα=jbk,α)(nkα=jβk,α))+nkj=1ϕk,j(nkα=jbk,α)(nkα=jβk,α)

    Note, that for i,j{1,,m+1} the index k is defined as the largest index of all subgraphs, the paths from the root to the i-th and to the j-th subgraph pass (see [12] for details). This index does not influence our computation, so we do not go into detail here.

    Proof of Lemma 3.5. For i{1,,m+1} and b,βM we have

    Σk,i(λb+(1λ)β)=i1k=k+11k1=kugk,nk(λbk+(1λ)βk)+gk,nk(λbk+(1λ)βk).

    Then from Lemma 3.1 it follows

    Σk,i(λb+(1λ)β)=i1k=k+11k1=ku(λ2gk,nk(bk)+(1λ)2gk,nk(βk)
    + 2(λλ2)nkj=1ϕk,j(nkα=jbk,α)(nkα=jβk,α))+(λ2gk,nk(bk)+(1λ)2gk,nk(βk)+ 2(λλ2)nkj=1ϕk,j(nkα=jbk,α)(nkα=1βk,α)),

    which can be written as

    Σk,i(λb+(1λ)β)=λ2((i1k=k+11k1=kugk,nk(bk))+gk,nk(bk))+ (1λ)2((i1k=k+11k1=kugk,nk(βk))+gk,nk(βk))+ 2(λλ2)((i1k=k+11k1=kunkj=1ϕk,j(nkα=jbk,α)(nkα=jβk,α))+nkj=1ϕk,j(nkα=jbk,α)(nkα=jβk,α)).

    With the definition of the sum in (19) it follows

    Σk,i(λb+(1λ)β)=λ2Σk,i(b)+(1λ)2Σk,i+2(λλ2)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 G=(V,E) with compressor edges. Let pressure bounds p+,min,p+,maxRn+1 with pmaxipminj (for all i=0,,n), constants ϕiRn0 (i=1,,n) and controls ui (for i=1,,m) be given. Additionally, let

    (pmaxi,0)2ui1(pmaxi1,0)2 (27)

    hold for i=2,,m+1. Then the set of feasible loads for a linear graph with compressor edges is convex.

    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 b+,β+M (let M be the set of feasible loads), then b,βRn0. For λ(0,1), (25) holds. Also the inequalities (8)-(10) hold for the load vector (λb+(1λ)β) for every subgraph because of Theorem 3.4. We have to show, that inequalities (11)-(18) hold for the convex combination (λb+(1λ)β). For indices i,j{1,,m+1} with i<j, the subgraph Gi is on the path from G1 to Gj, so the index k is equal to i and thus the sum Σk,i() is zero for the inequalities (11) - (18), but the sum Σk,j() not. Define

    c4,i,j:=1Πk,j(pmaxj,0)21Πk,i(pmini,0)2,

    with Πk,i defined in (20). Because k=i, it is Πk,i=1 and with (27), it follows

    1Πk,j(pmaxj,0)21Πk,juj1(pmaxj1,0)21Πk,j(j1k=kuk)(pmaxk,0)2.

    With the definition of Πk,j (see 20), we have

    1Πk,j(pmaxj,0)2(pmaxk,0)2.

    Thus we have c4,i,j0. We define

    t4,i,j(b,β)=c4,i,j+Σk,j(λb+(1λ)β).

    Then we can use Lemma 3.5 to get

    t4,i,j(b,β)=c4,i,j+λ2Σk,j(b)+(1λ)2Σk,j(β)+2(λλ2)Σk,j(b,β).

    Now we use inequality (11) itself for the feasible vectors b and β and get the estimate

    t4,i,j(b,β)c4,i,j+λ2(c4,i,j)+(1λ)2(c4,i,j)+2(λλ2)Σk,j(b,β),

    and from this it follows

    t4,i,j(b,β)2(λλ2)c4,i,j+2(λλ2)Σk,j(b,β).

    Because c4,i,j0 and Σk,j(b,β)0, it follows t4,i,j(b,β)0 and thus, inequality (11) holds for (λb+(1λ)β). For the next inequality, we define

    c5,i,j:=1Πk,i(pmaxi,0)21Πk,j(pmink,j)2,

    and

    t5,i,j(b,β):=c5,i,jΣk,j(λb+(1λ)β).

    Because Πk,i=1 and Πk,j1, we have c5,i,j0. We use Lemma 3.5 to get

    t5,i,j(b,β)=c5,i,j(λ2Σk,j(b)+(1λ)2Σk,j(β)+2(λλ2)Σk,j(b,β)).

    Then, we use the inequality (12) itself for feasible b and β and get

    t5,i,j(b,β)c5,i,j+λ2(c5,i,j)+(1λ)2(c5,i,j)2(λλ2)Σk,j(b,β),

    from which it follows

    t5,i,j(b,β)2(λλ2)c5,i,j2(λλ2)Σk,j(b,β).

    Now we use Lemma 3.6. It follows either

    t5,i,j(b,β)2(λλ2)c5,i,j2(λλ2)Σk,j(b),ort5,i,j(b,β)2(λλ2)c5,i,j2(λλ2)Σk,j(β).

    In both cases, we can use again inequality (12) to get

    t5,i,j(b,β)2(λλ2)c5,i,j2(λλ2)c5,i,j,

    which is obviously non-negative. Thus inequality (12) holds for (λb+(1λ)β). The proof for the inequalities (13), (15) and (17) works analogously to the proof of (11) and the proof of (14), (16) and (18) works analogously to the proof of (12). Then every inequality of Theorem 2.4 holds for the load vector (λb+(1λ)β) and thus the theorem is proven.

    Remark 2. The extra condition (pmaxi)2ui1(pmini1)2 of Theorem 3.7 is sufficient for convexity, but not necessary. The feasible set of a linear graph with compressor edges can also be convex even if the extra condition is not fulfilled.

    The following example illustrates the results.

    Example 3. : Consider the minimal graph with a compressor edge shown in Figure 9.

    Figure 9.  Linear graph with one compressor edge of example 3.

    We consider two different controls for pressure bounds (p+,min)=[1,1,1,1]T and (p+,max)=[3,2,2,1.5]T to show the results to show the results of Theorem 3.7. In the first case, we consider ua=2 and in the second case, we consider ub=4. Then from Theorem 2.4 we get

    Ma{bRR03|(b1+b2+b3)28b233(b1+b2+b3)2+12b238.75},Mb{bR30|(b1+b2+b3)28b233(b1+b2+b3)2+14b238.75(b1+b2+b3)2+14b230.4375}.

    The feasible sets are shown in Figure 10 and Figure 11.

    Figure 10.  Set Ma for (p+,min)a=[1,1,1,1]T and (p+,max)a=[3,3,3,3]T.
    Figure 11.  Set Mb for (p+,min)b=[2.5,2,1.5,1]T and (p+,max)b=[3,2.5,2,1.5]T.

    In both cases, one can see that the feasible sets are convex. Even if in case (b) the condition (pmaxi,0)2ui1(pmaxi1,0)2 is not fulfilled. But this is stated in Remark 2. The condition pmaxipminj is also not fulfilled in case (b), this leads to the case, that the vector 03Mb.

    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=(V,E) be a tree-structured graph without compressor edges. For bRn and λ(0,1), the pressure loss function defined in (6) is:

    g(λb)=λ2g(b). (28)

    Proof. Consider bRn and λ(0,1). We have

    g(λb)=(AT)1Φ|A1λb|(A1λb)=(AT)1Φ|λ||A1b|λ(A1b)=λ2(AT)1Φ||A1b|(A1b)=λ2g(b),

    since λ is non-negative.

    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 p+,min,p+,maxRn+1 with pmaxipminj (for all i,j=0,,n) be given. If the network graph is tree-structured with one input node and does not contain compressor edges, then the set of feasible loads M is star-shaped with respect to the point 0Rn+1.

    Proof. Let M be the set of feasible loads. To show this result, we have to show that for a feasible load vector bMRn0, the vector λb is also feasible for λ[0,1]. That means the vector λb fulfills the inequalities in Theorem 2.3 (for b=0, it is λb=0). First mention, that 0 is feasible. If b=0, it follows g(b)=0 and because pmaxipminj for all i,j=0,,n, it follows pmaxipminj0 for all i,j=0,,n. Thus, all inequalities in Theorem 2.3 hold and 0 is feasible. So if the set of feasible loads contains only one element, the statement of Theorem 4.2 is obviously true. Else, consider bM{0}. For i1,,n, we define

    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 b is feasible, we can use the first inequality in Theorem 2.3. It follows

    t1,i(b)c1,iλ2c1,i.

    Because c1,i0 (due to pmaxipminj) and λ[0,1], it follows t1,i(b)0 and thus the first inequality in Theorem 2.3 holds. Next define

    c2,i:=(pmax0)2(pmini)2,andt2,i(b):=c2,igi(λb).

    Then due to Lemma 4.1 we have

    t1,i(b)=c1,iλ2gi(b).

    We use the second inequality of Theorem 2.3 and get

    t1,i(b)c1,iλ2c1,i,

    and it follows t1,i(b)0. Thus, the second inequality in Theorem 2.3 holds. Last, we define

    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,jλ2c3,i,j,

    so it follows t3,i,j(b)0. Thus all inequalities of Theorem 2.3 are fulfilled for (λb) and the proof is complete.

    With the next example, we illustrate why we need the statement p+,maxp+,min.

    Example 4: Consider the minimal tree shown in Figure 12.

    Figure 12.  Graph of example 2.

    We consider the two cases (p+,min)a=[2,1,1]T, (p+,max)a=[3,2,2]T with the feasible set Ma) and (p+,min)b=[2.5,1.5,1]T, (p+,max)b=[3,2.5,2]T with the feasible set Mb). Then, for ϕ1=ϕ2, we get from Theorem 2.3

    Mb{bR20|b1,b28b213+b22b223+b21},Mb{bR20|b21.5b16.75b28b211.75+b22b225.25+b21},

    The feasible sets are shown in Figure 13.

    Figure 13.  Feasible sets for different pressure bounds.

    One can see, that in case (a), the set of feasible loads is star-shaped to the point 0R2, in case (b) this is not true. This is because the condition for the pressure bounds does not hold. But the set in case (b) still has the special property that the intersection of every line through the root and the feasible set is convex. This property is stated later in Lemma 4.4.

    Example 5: Consider the minimal tree shown in Figure 14.

    Figure 14.  Graph of example 5.

    We consider the two cases pmina=[1,1,1,1]T, pmaxa=[3,2,2,2]T with the feasible set Ma and pminb=[2,1,1,1]T, pmaxb=[3,2,2,1.5]T with the feasible set Mb. From Theorem 2.3 it follows

    Ma={bR30|b218(b2+b3)2+b238b213+(b2+b3)2(b2+b3)2+b233+b21b233},
    and     Mb={bR30|(b2+b3)2+b231.75b218(b2+b3)2+b238b213+(b2+b3)2b211.25+(b2+b3)2+b23(b2+b3)2+b233+b21b233}

    The feasible sets are shown in Figure 15 and Figure 16.

    Figure 15.  Set Ma for (p+,min)a=[1,1,1,1]T and (p+,max)a=[3,2,2,2]T.
    Figure 16.  Set Ma for (p+,min)b=[2,1,1,1]T and (p+,max)b=[3,2,2,1.5]T.

    As in Example 4, the feasible set in case (a) is star-shaped to the point 0R3, the set in case (b) is not. But again, the intersection of every line through the root and the set of feasible loads is convex (see Lemma 4.4).

    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 uiR (i=1,,m) be given. Let pmaxipminj hold for every subgraph and let (27) hold, i.e. (pmaxi,0)2ui1(pmaxi1,0)2. Then the set of feasible loads of a tree-structured graph with compressor edges is star-shaped to the point 0Rn.

    Proof. Consider b+Rn+1 and λ[0,1]. We have to show, that the inequalities (8)-(18) hold for the load vector λb. Theorem 4.2 we know, that (8)-(10) are fulfilled for every subgraph. For the sum defined in (19) we know from Lemma 4.1 (i{1,,n})

    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)21Πk,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)21Πk,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 b is feasible, we can use inequality (11) resp. (12) to get the estimates

    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 c4,i,j,c5,i,j0 because of the condition (pmaxi,0)2ui1(pmaxi1,0)2. So both, t4,i,j(b) and t5,i,j(b) are non-negative due to our assumptions. Thus inequality (11) and (12) hold for (λb). All other inequalities can be shown analogously. That means the feasible set for tree-structured graphs with compressor edges is star-shaped to the point 0n+1 and thus the proof is complete.

    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 MRn be the set of feasible loads of a tree-structured graph. Then, for a point bRn, the set L:={βM|β=λb (λ[0,1])} is convex.

    Remark 3. The statement of Lemma 4.4 is a generalized star-shapedness property with respect to the point 0Rn. A set S is star-shaped with respect to a point s if sS and if for every direction dRn, the line from s in direction d has a convex intersection with the set S. If S is generalized star-shaped to the point s, the same property holds for sS, so here the point s need not to be in the set S. For the computation in the spheric-radial decomposition, this situation is as useful as the classical star-shapedness property.

    Proof of Lemma 4.4. First, if the set L is empty or contains only one element, it is convex. We only consider the case bRn0, because otherwise, the set L contains at most one element (the point 0Rn). This is, because all load vectors (without the first component) are non-negative due to our network graph has only one inflow node. We separate the proof in two parts. In the first part, we prove Lemma 4.4 for tree-structured networks without compressor edges, in the second part we prove it for general tree-structured networks.

    Part I: Proof for trees without compressor edges: We first prove this lemma for trees without compressor edges. Consider bRn0 and assume, that the feasible set contains at least two elements. Consider β1,β2M with β1=λ1b and β2=λ2b (0λ1<λ21). Then, the inequalities in Theorem 2.3 hold for β1 and β2. We have to show, that these inequalities also hold for β=λb for all λ[λ1,λ2]. The first inequality for β1 is

    (pmin0)2(pmaxk)2gk(β1),

    which is equal to

    (pmin0)2(pmaxk)2gk(λ1b).

    From Lemma 4.1, it follows

    (pmin0)2(pmaxk)2λ21gk(b),

    and this also holds for every λλ1, especially for λ[λ1,λ2]. The second inequality for β2 is

    (pmax0)2(pmin0)2gk(β2).

    This is equal to

    (pmax0)2(pmin0)2gk(λ2b),

    and due to Lemma 4.1 it follows

    (pmax0)2(pmin0)2λ22gk(b).

    This also holds for λλ2, especially for every λ[λ1,λ2]. The third inequality in Theorem 2.3 is

    (pmink)2(pmax)2gk(β1)g(β1),resp.(pmink)2(pmax)2gk(β2)g(β2).

    The term on the right in both cases has the same sign, because if gk(β1)g(β1)<0, then due to Lemma 4.1 this is equal to λ21(gk(b)g(b))<0 and the sign is independent of λ1. Thus this also holds for β2=λ2b. So if gk(b)g(b)0, we follow the argumentation of the first inequality, we have shown here. And if gk(b)g(b)<0, we follow the argumentation of the second inequality, we have shown here. Thus, all inequalities in Theorem 2.3 hold for β=λb for all λ[λ1,λ2] and the lemma is proven for trees without compressor stations.

    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 bRn0 and β1,β2M with β1=λ1b and β2=λ2b (0λ1<λ21). We have to show, that all inequalities in Theorem 2.4 hold for β=λb with λ[λ1,λ2]. The inequalities (8) - (10) follow directly from the first part of the proof. Consider inequality (11):

    1Πk,i(pmini,0)21Πk,j(pmaxj,0)2Σkj(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 β=λb with λ[λ1,λ2]. Thus, all inequalities in Theorem 2.4 are fulfilled for β=λb with λ[λ1,λ2] and the set L is convex. So Lemma 4.4 is proven.

    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 0Rn. For even weaker assumptions, we introduced the generalized star-shapedness (see Lemma 4.4), which is also very useful for the computation in the spheric-radial decomposition.

    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 0Rn. Convexity in this case does not hold, not even in the simplest case of a network cycle of three nodes (see [9], {Section 5}).

    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.
  • This article has been cited by:

    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
  • Reader Comments
  • © 2020 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(2279) PDF downloads(343) Cited by(5)

Figures and Tables

Figures(16)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog