Coupling conditions for the transition from supersonic to subsonic fluid states

  • Received: 01 May 2016 Revised: 01 July 2017
  • Primary: 35L60, 35L04; Secondary: 35R02

  • We discuss coupling conditions for the p-system in case of a transition from supersonic states to subsonic states. A single junction with adjacent pipes is considered where on each pipe the gas flow is governed by a general p-system. By extending the notion of demand and supply known from traffic flow analysis we obtain a constructive existence result of solutions compatible with the introduced conditions.

    Citation: Martin Gugat, Michael Herty, Siegfried Müller. Coupling conditions for the transition from supersonic to subsonic fluid states[J]. Networks and Heterogeneous Media, 2017, 12(3): 371-380. doi: 10.3934/nhm.2017016

    Related Papers:

    [1] Martin Gugat, Michael Herty, Siegfried Müller . Coupling conditions for the transition from supersonic to subsonic fluid states. Networks and Heterogeneous Media, 2017, 12(3): 371-380. doi: 10.3934/nhm.2017016
    [2] 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
    [3] Yannick Holle, Michael Herty, Michael Westdickenberg . New coupling conditions for isentropic flow on networks. Networks and Heterogeneous Media, 2020, 15(4): 605-631. doi: 10.3934/nhm.2020016
    [4] Christophe Chalons . Theoretical and numerical aspects of the interfacial coupling: The scalar Riemann problem and an application to multiphase flows. Networks and Heterogeneous Media, 2010, 5(3): 507-524. doi: 10.3934/nhm.2010.5.507
    [5] Rinaldo M. Colombo, Francesca Marcellini . Coupling conditions for the $3\times 3$ Euler system. Networks and Heterogeneous Media, 2010, 5(4): 675-690. doi: 10.3934/nhm.2010.5.675
    [6] Steinar Evje, Kenneth H. Karlsen . Hyperbolic-elliptic models for well-reservoir flow. Networks and Heterogeneous Media, 2006, 1(4): 639-673. doi: 10.3934/nhm.2006.1.639
    [7] Raul Borsche, Axel Klar, T. N. Ha Pham . Nonlinear flux-limited models for chemotaxis on networks. Networks and Heterogeneous Media, 2017, 12(3): 381-401. doi: 10.3934/nhm.2017017
    [8] Yogiraj Mantri, Michael Herty, Sebastian Noelle . Well-balanced scheme for gas-flow in pipeline networks. Networks and Heterogeneous Media, 2019, 14(4): 659-676. doi: 10.3934/nhm.2019026
    [9] Rinaldo M. Colombo, Graziano Guerra . A coupling between a non--linear 1D compressible--incompressible limit and the 1D $p$--system in the non smooth case. Networks and Heterogeneous Media, 2016, 11(2): 313-330. doi: 10.3934/nhm.2016.11.313
    [10] Jean-Marc Hérard, Olivier Hurisse . Some attempts to couple distinct fluid models. Networks and Heterogeneous Media, 2010, 5(3): 649-660. doi: 10.3934/nhm.2010.5.649
  • We discuss coupling conditions for the p-system in case of a transition from supersonic states to subsonic states. A single junction with adjacent pipes is considered where on each pipe the gas flow is governed by a general p-system. By extending the notion of demand and supply known from traffic flow analysis we obtain a constructive existence result of solutions compatible with the introduced conditions.



    Natural gas pipeline transportation systems are an important part of our infrastructure. Besides transport of gas, also CO2 pipeline transport plays an important role in carbon capture and storage, see for example [18]. A well-established mathematical model for such flow is the p-system (1). The p-system in conservative variables density ρ>0 and flux q is given by

    tρ(t,x)+xq(t,x)=0,tq(t,x)+x(q2(t,x)ρ(t,x)+p(ρ(t,x)))=0, (1)

    where t0 is time, x is space, and ρp(ρ) is the pressure law. The system is stated in Eulerian coordinates and it is a simplified model for gas flow in pipes [21] or water flow in open canals [11]. In the latter case ρ=h is the water height and p(h)=g2h2 with g denoting the gravitational acceleration. We are interested in the situation of a single junction connected to pipes depicted in Figure 1. The pipes are modeled as directed arcs and the junction as node. We suppose to have n incoming pipes parameterized by x(,0] and labeled by kδ:={n,,1}. Similarly, we suppose to have m outgoing pipes kδ+:={1,,m} and parameterized by x[0,). Along each pipe the evolution of the gas is supposed to fulfill (1) and the state of the system in pipe k is denoted by Uk:=(ρk,qk)t. The dynamics of the pipes is then coupled at x=0 through coupling conditions [2]. General results on the p-system at a junction are given in [9]. For a recent overview of theoretical and numerical results for flows on networks governed by hyperbolic equations we refer to [4]. Except for [14] all current coupling conditions rely on a transversality condition [10] to establish well-posedness. This condition is replaced here by a solvability condition prior to Theorem 2.3 having a similar meaning, namely, that Lax curves are not parallel in the point of intersection in phase space.

    Figure 1.  Junction of n+m=|δ|+|δ+| connected pipes.

    It has been shown that the transversality condition is fulfilled for subsonic data, i.e. initial conditions U0i(x) with iδ+δ fulfilling λ1(U0i)<0<λ2(U0i) where λj,j=1,2 are the characteristic velocities given by equation (5). To the best of our knowledge, no results in the supersonic case exist. However, in the scalar case on networks this case has been well-understood and discussed intensively, see for example [17,6] or [12]. We are interested in the coupling of supersonic and subsonic states for the p-system. We focus on the most common coupling condition for gas networks as well as coupled open canals, namely, we prescribe

    p(ρk(t,0+))=p(t)kδ+,p(ρk(t,0))=p(t)kδ, (2a)
    kδqk(t,0)=kδ+qk(t,0+). (2b)

    for some unknown nodal pressure p(t) and where 0± denote the trace values coming from the interior of the pipe. We refer to [2,15,9] for a more detailed discussion of coupling conditions. Our main result is Theorem 2.3 showing that there exists a solution to the coupling conditions (2) under suitable assumptions on the initial data. The key assumption is on the sign eigenvalues of the initial datum, i.e., equation (17). Briefly summarized the assumption states that the data on incoming arcs is supersonic and the data on outgoing arcs is subsonic.

    The study of the coupling of subsonic and supersonic flow is also important for algorithmic purposes. For example in an iterative method for the solution of optimal control problems (see for example [7]), it can happen that during the iteration in an intermediate step, supersonic states appear even if the optimal state is subsonic. The transition from supersonic to subsonic flow is important for high-speed aircraft, where a supersonic external freestream flow must be decelerated efficiently to subsonic speed at the turbine inlet. This transition from supercritical flow to subcritical flow also occurs in Fanno flows, see [22], (p. 199). The transition from supercritical to subcritical regime in free surface flow of yield stress fluids has been studied in [23] where a hydraulic jump during laminar flow of an Herschel-Bukley fluid in a rectangular channel is studied. An experimental assessment of scale effects affecting two-phase flow properties in hydraulic jumps is given in [20].

    In this work we will introduce a notion of supply and demand for the state of the system on the adjacent pipes similar to the notion in traffic flow [19,14]. This will allow to solve for constant initial data U0k=cst the problem (1) and (2) also in the supersonic case. Necessary conditions on the constants U0k are presented ensuring solvability. In the case of n=m=1 the solution coincides with a classical solution to a Riemann problem.

    We recall the Lax-curves for the p-system for a pressure law satisfying the following assumption on the equation of state, i.e. pC3(R+;R+) with

    p(ρ)>0, p(ρ)0, (3a)
    6p(ρ)+3p(ρ)(ρˉρ)+ρ(3p(ρ)+(ρˉρ)p(ρ))>0ρ>ˉρ0 (3b)

    and such that p(ρ),p(ρ)0 for ρ0. Note that the first condition (3a) implies the usual assumption on the equation of state

    ρp(ρ)+2p(ρ)>0, (4)

    see e.g. [9,5]. This also implies that both characteristic fields are genuinely nonlinear. In addition, the second condition (3b) is required later in order to obtain strict convex (resp. concave) wave curves. The assumptions (3) are fulfilled by the pressure law of the isentropic Euler equations p(ρ)=ργ for γ>1 as well as for the shallow-water equations where the hydrostatic pressure is p(ρ)=g2ρ2. In particular, the isothermal equation p(ρ)=ρ is frequently applied in network simulations. We note that a sufficient condition for (3b) is p(ρ)0 that for isentropic Euler equations only holds for γ2.

    Under the assumption (3a) the p-system (1) is strictly hyperbolic in the non-vacuum states {U:ρ>0}. The speed of sound is given by c(ρ)=p(ρ) and the eigenvalues of the ith characteristic field are for i=1,2

    λ1(U)=qρc(ρ), λ2(U)=qρ+c(ρ). (5)

    Both characteristic fields are genuinely nonlinear if the equation of state fulfills (4) and (3a). The 1-Lax-curve through ˉU=U is given by

    L1(σ;ˉU)=(R1(σ;ˉU)σˉρS1(σ;ˉU)σ>ˉρ). (6)

    The reversed 2-Lax-curve through ˉU=Ur is given by

    L2(σ;ˉU)=(S2(σ;ˉU)σ>ˉρR2(σ;ˉU)σˉρ). (7)

    Here, for i=1,2 we have the positive and negative branch of

    Si(σ;ˉU)=σ ˉqˉρσˉρ(σˉρ)(p(σ)p(ˉρ)),Ri(σ;ˉU)=σˉqˉρσσˉρc(s)sds.

    The precise form and derivation of the curves can be found e.g. in [9,13]. Therein, also the following properties are established: The shock curves Si,i=1,2, are concave and convex with respect to U in the ρ-q-plane. The rarefaction curves Ri,i=1,2, are concave and convex, respectively, in the ρ-q-plane. The Riemann problem has a global unique solution, see [5,Section 6.3]. The remaining Lax curves are given for example in [9,Equation 4.1,Equation 4.2].

    From the explicit form of Ri we obtain

    d2dσ2Ri(σ,ˉU)=(σp(σ)+2p(σ)2σp(σ)).

    Since σ>ˉρ>0, the discriminant s(σ)=σˉρ(σˉρ)(p(σ)p(ˉρ)) in Si(σ;ˉU) is positive. Furthermore, due to assumption (3), s is strict convex, i.e.

    s(σ)=2σˉρp(σ)+2ˉρ(p(σ)p(ˉρ))+σˉρˉρ(2p(σ)+σp(σ))>0.

    Hence,

    d2dσ2Si(σ,ˉU)=12s(σ)(s(σ)s(σ)22s(σ)).

    The function ˜s(σ)=2s(σ)s(σ)s(σ)2 vanishes at σ=ˉρ and ˜s(σ)=2s(σ)s(σ).

    We also note that

    s(σ)ˉρ=6p(σ)+3p(σ)(σˉρ)+σ(3p(σ)+(σˉρ)p(σ))>0

    provided that condition (3b) holds true. Then, ˜s is a monotone increasing function for all σ>ˉρ starting at zero. This yields the strict concavity and convexity of Si, respectively.

    Under assumption (3b) we therefore obtain

    Si is strictly concave and Ri is strictly convex,i=1,2. (8)

    Hence, L1 and L2 have a single maximum and minimum, respectively. We denote by q=q(ˉU) the maximum of the functions ρL1(ρ;ˉU) and ρL2(ρ;ˉU). The corresponding density is denoted by ρ=ρ(ˉU). The {Heaviside} function is denoted by H(x).

    We now define the demand and supply for the p-system extending the notion in the scalar case [19]. In the ρ-q-plane and for ρ>0 the demand ρd(ρ;ˉU) is an extension of the non-decreasing part of the curve ρL1(ρ;U). In the ρ-q-plane and for ρ>0 the supply function ρs(ρ;ˉU) is an extension of the non-increasing part of the curve ρL2(ρ;U). The maximum of L1 and L2 depend on ˉU. For readability we skip this dependence in the following formulas for d and s.

    d(ρ;ˉU)=(L1(ρ;ˉU),0<ρmin{ˉρ,ρ}ˉqH(ρˉρ)+qH(ˉρρ), else ) (9)
    s(ρ;ˉU)=(L2(ρ;ˉU)H(ρρ)qH(ρρ),ˉρ>ρˉqH(ˉρρ)+H(ρˉρ)max{ˉq,L2(ρ;ˉU)},ˉρ<ρ). (10)

    A sketch of the supply and demand for different states ˉU=(ˉρ,ˉq) is given in Figure 2 and Figure 3, respectively.

    Figure 2.  The supply function ρs(ρ;ˉU) in red for given data ˉU indicated by a cross for ˉρ>ρ (left) and ˉρ<ρ (right). Also shown in blue is the curve L2. The state U is indicated by a circle..
    Figure 3.  The demand function ρd(ρ;ˉU) in red for given data ˉU. Also shown in blue is the curve L1..

    The demand and supply are used to determine admissible fluxes for waves of non-positive and non-negative speed, respectively. A Riemann problem is a Cauchy problem for equation (1) and initial data U0(x)=(Ux<0Urx>0). A weak solution to the Riemann problem is defined e.g. in [3] and consists of a superposition of self-similar waves. In the case of the 2×2 system we obtain at most two waves separated by an intermediate state. The general structure in the present case therefore is a superposition of either a shock or rarefaction wave of the first family with a shock or rarefaction wave of the second family.

    Theorem 2.1. Let ˉU with ˉρ>0 be given and assume (3). Then, for all qd(ˉρ;ˉU), there exists a state ˜U=(˜ρ,q) such that the Riemann problem for (1) with U=ˉU and Ur=˜U admits a weak self-similar solution composed of waves of non-positive speed.

    Proof. We have to distinguish two cases. Suppose ˉρ<ρ(ˉU), see Figure 3 (right). Then, d(ˉρ;ˉU)=ˉq. Denote by τˉρ the value of ρ, such that ˉq=L1(τ;ˉU). Due to the strict concavity of L1 and for q<d(ˉρ;ˉU) there exists a unique density ˜ρ>τ, such that L1(˜ρ;ˉU)=q. The solution to the Riemann problem consists of a single shock propagating with non-positive velocity. Suppose ˉρ<ρ(ˉU) and assume q=ˉq=d(ˉρ;ˉU). Then, ˜U=ˉU as well as ˜U=(τ,q) where τˉρ fulfills L1(τ;ˉU)=q. Note that in the latter case the solution is a shock of zero speed. Since this is a stationary discontinuity located at a single point we can exclude this solution from our construction as in [16].

    In the case ˉρ>ρ(ˉU), see Figure 3 (left), we obtain d(ˉρ;ˉU)=q(ˉU). For ˉqqq(ˉU) we find a density ˜ρ such that ρ˜ρ(ˉU)ˉρ and q=L1(˜ρ;ˉU). The solution to the corresponding Riemann problem consists of a rarefaction wave with non-positive speed. For q<ˉq, we find ˜ρ>ˉρ and q=L1(˜ρ;ˉU), such that the solution to the corresponding Riemann problem consists of a shock of negative speed.

    Corollary 1. Let ˉU with ˉρ>0 be given and assume (3). Then, for allqs(ˉρ;ˉU), there exists a state ˜U=(˜ρ,q) such that the Riemann problem for (1) withU=˜U and Ur=ˉU admits a weak self-similar solution composed of waves of non-negative speed.

    The proof is similar to the proof of Theorem 2.1 and omitted. It should be noted that in the formulation of Corollary 1 the right state ˉU is given instead of the left state in Theorem 2.1.

    We introduce the following subsets of R+×R depending on the sign of the eigenvalues λi given by (5).

    A+:={U:=(ρ,q):λi(U)>0,i=1,2}, (11)
    A:={U:=(ρ,q):λ1(U)<0<λ2(U)},A:={U:=(ρ,q):λi(U)<0,i=1,2}. (12)

    The set A comprises of the states in the subsonic region, A+A is the supersonic region. For the junction located at x=0 we consider the Riemann problem

    t(ρk(t,x)qk(t,x))+x(qk(t,x)q2k(t,x)ρk(t,x)+p(ρk(t,x)))=0,Uk(t,0)=(Ukx<0U+kx>0). (13)

    Depending on the pipe only one of the Riemann data is defined at t=0, i.e. Uk=U0,k if kδ and U+k=U0,k if kδ+. For each pipe the remaining unknown state U±k for kδ± is determined in such a way that equation (2) holds true. We construct a Riemann solver at the node (Uk(t,x))kδ± to (13) and (2). An explicit construction is provided for constant initial data self-similar solutions Uk(t,x) consisting of wave of non-positive speed and non-negative speed for kδ and kδ+, respectively. This solution is restricted to Uk(t,x) for x<0 and x>0, when kδ and kδ+, respectively.

    The demand and supply allows to construct solutions of certain wave speeds. Consider kδ. Let Uk:=U0,k and qd(ρ0,k;U0,k) be given. Then, according to Theorem 2.1, there exist a state U+k with q+k=q and such that the Riemann problem (13) admits a solution Uk(t,x). The solution consists of a wave of non-positive speed. Similarly, for kδ+, data U+k:=U0,k and any qs(ρ0,k;U0,k) a self-similar solution Uk(t,x) consisting of waves of non-negative speed exist according to Corollary 1.

    In order to obtain a unique density at the node we define the function ˉUτk(ˉU) by

    kδ:τk(ˉU):=(ρ(ˉU) if ˉρρ(ˉU),˜ρ else), where ˜ρˉρ and L1(˜ρ;ˉU)=ˉq. (14a)

    Similarly,

    kδ+:τk(ˉU)=(ρ(ˉU) if ˉρρ(ˉU),˜ρ else), where ˜ρˉρ and L2(˜ρ;ˉU)=ˉq. (14b)

    Clearly, τk>0 if ˉρ>0. τk is well-defined due to equation (8). For kδ and ˉρ<ρ(ˉU) we obtain L1(τk(ˉU);ˉU)=L1(ˉρ;ˉU)=ˉq. Hence, in this case τk(ˉU) is the density such that the wave connecting U=(ˉρ,ˉq) and Ur=(τk(ˉU),ˉq) has zero speed and is omitted as in the proof of Theorem 2.1.

    Finally, we turn to the description of the Riemann solver for problem (13) and (2). Due to assumption (3) we have that the condition (2) is equivalent to

    ρk(t,0+)=ρk(t,0)kδ+,kδ, (15)

    and

    kδqk(t,0)kδ+qk(t,0+)=0. (16)

    Note that in the case n=m=1, i.e. δ±={±1}, the conditions (15) and (2) imply by means of the Rankine-Hugoniot jump conditions in particular

    q1(t,0)=q1(t,0+),q21(t,0)ρ1(t,0)+p(ρ1(t,0))=q21(t,0+)ρ1(t,0+)+p(ρ1(t,0+)).

    For consistency a solution to (13) should therefore be also a solution to a (classical) Riemann problem for (1) with initial data U(0,x)=(U1x<0U+1x>0).

    We are interested in supersonic and subsonic coupling at the junction. Therefore, we assume in the following that

    Uk=U0,kA+ for kδ,U+k=U0,kA for kδ+. (17)

    Remark 1. Due to the symmetry of the Lax curves for the p-system, the discussion also includes the case where UkA for kδ and U+kA for kδ+. The case U±kA for kδ± has been discussed for example in [8].

    We now plan to construct a solution fulfilling the coupling conditions. As preliminary result we prove the monotonicity of the Lax curves.

    Lemma 2.2. Assume U0,k for kδ± with ρ0,k>0 are given. Suppose further that (3) holds true.

    Then, for all σσlb we have that σL1(σ;U0,k) for kδ is strictly monotone decreasing and σL2(σ;U0,k) for kδ+ is strictly monotone increasing.

    Proof. In the case ρ0,k<ρ(U0,k) we have that τk(U0,k)>ρ(U0,k) by definition (14). Therefore, the proof follows from the monotonicity properties of L1() and L2(), respectively. The monotonicity implies that σkδL1(σ;U0,k) and σkδ+L2(σ;U0,k) are monotone decreasing and increasing in [σlb,), respectively. Furthermore, L1(σlb;U0,k)d(σlb;U0,k),kδ and L2(σlb;U0,k)s(σlb;U0,k),kδ+. This finishes the proof.

    In general, under assumption (17), the solution to a Riemann problem for (1) consists of a superposition of waves of the first family and waves of the second family. In order to solve the problem (13) we mimic this behavior by constructing of a solution as follows. Denote by

    σlb:=maxkδ±τk(U0,k). (18)

    Then, equation (16) is fulfilled provided that there exists σ with σσlb and (16)

    kδL1(σ;U0,k)kδ+L2(σ;U0,k)=0. (19)

    The following Theorem guarantees the solvability of the previous equation and presents the main result.

    Theorem 2.3. Assume U0,k for kδ± with ρ0,k>0 are given. Suppose further that (3) holds true. Let σlb be given by equation (18).

    Provided that

    kδL1(σlb;U0,k)kδ+L2(σlb;U0,k) (20)

    there exists a unique σσlb, such that equation (19) is fulfilled.

    The proof follows immediately from the monotonicity properties established before.

    Now, we continue our construction: Suppose (19) has a unique solution ¯σ, see Theorem 2.3. Then, define

    U+k:=(¯σ,L1(¯σ;U0,k))T,kδ,Uk:=(¯σ,L2(¯σ;U0,k))T,kδ+. (21)

    Then, we have L1(¯σ;U0,k)d(ρ0,k;U0,k) for kδ and L2(¯σ;U0,k)s(ρ0,k;U0,k) for kδ+ due to Lemma 2.2. According to Theorem 2.1 and Corollary 1 this ensures that the solution to problem (13) has non-positive and non-negative speed, respectively. Therefore, a solution to the problem (13) fulfills equation (16) and (15).

    Some remarks are in order.

    Remark 2. Note that this construction is also valid in the case UkA and coincides with the solution to the equal pressure condition proposed in [8,2].

    In the case n=m=1 we have U+1=U1=:ˉU. Since λ1(ˉU)<λ2(ˉU) the solution (13) consists by construction of at most a wave of the first family on k=1. On k=1 we may have at most a wave of the second family. This is consistent with the solution to a Riemann problem (13) with initial data U(0,x)=(U1x<0U+1x>0).

    Remark 3. Note that linearization techniques in order to solve (2) and (16) will fail to provide a solution to (13) due to the fact that at U0,k,kδ both eigenvalues are non-negative. Therefore, a linear system obtained through linearization at U0,k,kδ± would have only m degrees of freedom to prescribe boundary conditions. However, the given set of conditions are in fact n+m. This is different compared with the case U0,kA for all kδ±.

    In the case U0,kA for kδ+ the problem (13), (15) and (16) is more complicated in the following way. The characteristic families can no longer be uniquely related to whether they are incoming (kδ) or outgoing (kδ+). One can easily find examples such that we observe two waves of different families on the outgoing and a wave on the incoming pipe. Therefore, new additional conditions have to be imposed to treat the additional degree of freedom.

    We consider p(ρ)=ρ2 as an example. Further, we assume δ={3,,1} and δ+={1,,4}. Initial data U0,k is chosen randomly such that U0,kA+ for kδ and U0,kA for kδ+. The initial data U0,k does not fulfill the coupling condition (16) and (15) as seen in Table 1. We then compute the states at the junction states at junction (21) according to the discussion of the previous section. Numerically, we solve equation (19) using Newton's method. The results are given in Table 2. As expected the density and the pressure fulfill (2) and (15). Further, we report on the difference between incoming and outgoing fluxes. The equation (19) has been solved up to machine precision.

    Table 1.  Random initial states U0,k on incoming pipes k<0 and outgoing pipes k>0. The initial difference in the sum of the mass fluxes is given by Δ=9.162e+00..
    Pipe k(ρ0,k,q0,k)p(ρ0,k)
    -1(5.151e-01, 2.519e+00)2.653e-01
    -2(6.317e-01, 2.794e+00)3.991e-01
    -3(6.642e-01, 3.905e+00)4.412e-01
    1(5.730e-01, -2.648e-01)3.283e-01
    2(7.460e-01, -1.523e-01)5.565e-01
    3(5.931e-01, -1.280e-01)3.518e-01
    4(5.849e-01, 6.020e-01)3.421e-01

     | Show Table
    DownLoad: CSV
    Table 2.  Terminal states Uk(t,0±) for kδ±. The difference in the sum of the mass fluxes is zero.
    Pipe k(ρk,qk)p(ρk)
    -1(2.089e+00, 5.101e+00)4.364e+00
    -2(2.089e+00, 4.868e+00)4.364e+00
    -3(2.089e+00, 8.090e+00)4.364e+00
    1(2.089e+00, 3.757e+00)4.364e+00
    2(2.089e+00, 3.357e+00)4.364e+00
    3(2.089e+00, 4.147e+00)4.364e+00
    4(2.089e+00, 6.798e+00)4.364e+00

     | Show Table
    DownLoad: CSV

    While coupling conditions for the subsonic flow in gas pipelines are well-established, in this paper the coupling conditions for the transonic transition from supersonic flow to subsonic flow are studied that have not been thoroughly analyzed before. In this way we fill a gap in the theory, which is important in order to have a complete theory that covers all cases. This can also be important for algorithmic purposes, since during an iterative method for the solution of an optimal control problem, it may happen that supersonic states occur, even if the optimal state is subsonic. The purely subsonic case has been treated explicitly in [1] and in general e.g. in [8]. For a complete theory of all possible cases the fully supersonic case is still at large. Future work may then include a complete statement provided the challenges in the supersonic regime can be dealt with.

    [1] Coupling conditions for gas networks governed by the isothermal Euler equations. Netw. Heterog. Media (2006) 1: 295-314 (electronic).
    [2] Gas flow in pipeline networks. Netw. Heterog. Media (2006) 1: 41-56.
    [3] A. Bressan, Hyperbolic Systems of Conservation Laws, The One-Dimensional Cauchy Problem, Oxford Lecture Series in Mathematics and its Applications, 20, Oxford University Press, Oxford, 2000.
    [4] Flow on networks: recent results and perspectives. European Mathematical Society-Surveys in Mathematical Sciences (2014) 1: 47-111.
    [5] The Cauchy problem for the Euler equations for compressible fluids. Handbook of Mathematical Fluid Dynamics (2002) 1: 421-543.
    [6] Traffic flow on a road network. SIAM J. Math. Anal. (2005) 36: 1862-1886 (electronic).
    [7] Optimal control in networks of pipes and canals. SIAM J. Control Optim. (2009) 48: 2032-2050.
    [8] A well posed Riemann problem for the p-system at a junction. Netw. Heterog. Media (2006) 1: 495-511.
    [9] On the Cauchy problem for the p-system at a junction. SIAM J. Math. Anal. (2008) 39: 1456-1471.
    [10] On 2×2 conservation laws at a junction. SIAM J. Math. Anal. (2008) 40: 605-622.
    [11] Thèorie du mouvement non-permanent des eaux, avec application aux crues des rivière at à l'introduction des marèes dans leur lit.. C.R. Acad. Sci. Paris (1871) 73: 147-154.
    [12] M. Garavello and B. Piccoli, Traffic Flow on Networks, vol. 1 of AIMS Series on Applied Mathematics, American Institute of Mathematical Sciences (AIMS), Springfield, MO, 2006, Conservation laws models.
    [13] E. Godlewski and P. -A. Raviart, Numerical Approximation of Hyperbolic Systems of Conservation Laws, Applied Mathematical Sciences, 118, Springer-Verlag, New York, 1996. doi: 10.1007/978-1-4612-0713-9
    [14] Coupling conditions for a class of second-order models for traffic flow. SIAM J. Math. Anal. (2006) 38: 595-616.
    [15] Assessment of coupling conditions in water way intersections. Internat. J. Numer. Methods Fluids (2013) 71: 1438-1460.
    [16] A mathematical model of traffic flow on a network of unidirectional roads. SIAM J. Math. Anal. (1995) 26: 999-1017.
    [17] Riemann problems with a kink. SIAM J. Math. Anal. (1999) 30: 497-515 (electronic).
    [18] S. Joana, M. Joris and T. Evangelos, Technical and Economical Characteristics of Co2 Transmission Pipeline Infrastructure, Technical report, JRC Scientic and Technical Reports, European Commission.
    [19] Les modeles macroscopiques du traffic. Annales des Ponts. (1993) 67: 24-45.
    [20] Experimental assessment of scale effects affecting two-phase flow properties in hydraulic jumps. Experiments in Fluids (2008) 45: 513-521.
    [21] Simulation of transient flow in gas networks. Int. Journal for Numerical Methods in Fluid Dynamics (1984) 4: 13-23.
    [22] B. Sultanian, Fluid Mechanics: An Intermediate Approach, CRC Press, 2015.
    [23] R. Ugarelli and V. D. Federico, Transition from supercritical to subcritical regime in free surface flow of yield stress fluids Geophys. Res. Lett. , 34 (2007), L21402. doi: 10.1029/2007GL031487
  • This article has been cited by:

    1. E. Fokken, S. Göttlich, O. Kolb, Modeling and simulation of gas networks coupled to power grids, 2019, 119, 0022-0833, 217, 10.1007/s10665-019-10026-6
    2. EIKE FOKKEN, SIMONE GÖTTLICH, MICHAEL HERTY, Efficient simulation of coupled gas and power networks under uncertain demands, 2022, 0956-7925, 1, 10.1017/S0956792522000079
    3. Andrea Corli, Ulrich Razafison, Massimiliano D. Rosini, Coherence and flow-maximization of a one-way valve, 2022, 56, 2822-7840, 1715, 10.1051/m2an/2022053
    4. Georges Bastin, Jean-Michel Coron, Amaury Hayat, Peipei Shang, Boundary feedback stabilization of hydraulic jumps, 2019, 7, 24686018, 100026, 10.1016/j.ifacsc.2019.100026
    5. Michael Herty, Siegfried Müller, Aleksey Sikstel, Coupling of Compressible Euler Equations, 2019, 47, 2305-221X, 769, 10.1007/s10013-019-00353-7
    6. Martin Gugat, Michael Herty, 2022, 23, 9780323850599, 59, 10.1016/bs.hna.2021.12.002
    7. Michael Herty, Niklas Kolbe, Siegfried Müller, Central schemes for networked scalar conservation laws, 2022, 18, 1556-1801, 310, 10.3934/nhm.2023012
    8. 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
  • © 2017 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(4363) PDF downloads(205) Cited by(8)

Figures and Tables

Figures(3)  /  Tables(2)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog