
In dynamical systems on networks, Kirchhoff's first law describes the local conservation of a quantity across edges. Predominantly, Kirchhoff's first law has been conceived as a phenomenological law of continuum physics. We establish its algebraic form as a property that is inherited from fundamental axioms of a network's geometry, instead of a law observed in physical nature. To this end, we extend calculus to networks, modeled as abstract metric spaces, and derive Kirchhoff's first law for hyperbolic conservation laws. In particular, our results show that hyperbolic conservation laws on networks can be stated without explicit Kirchhoff-type boundary conditions.
Citation: Alexandre M. Bayen, Alexander Keimer, Nils Müller. A proof of Kirchhoff's first law for hyperbolic conservation laws on networks[J]. Networks and Heterogeneous Media, 2023, 18(4): 1799-1819. doi: 10.3934/nhm.2023078
[1] | Martin Gugat, Alexander Keimer, Günter Leugering, Zhiqiang Wang . Analysis of a system of nonlocal conservation laws for multi-commodity flow on networks. Networks and Heterogeneous Media, 2015, 10(4): 749-785. doi: 10.3934/nhm.2015.10.749 |
[2] | Mauro Garavello . A review of conservation laws on networks. Networks and Heterogeneous Media, 2010, 5(3): 565-581. doi: 10.3934/nhm.2010.5.565 |
[3] | Gabriella Bretti, Roberto Natalini, Benedetto Piccoli . Numerical approximations of a traffic flow model on networks. Networks and Heterogeneous Media, 2006, 1(1): 57-84. doi: 10.3934/nhm.2006.1.57 |
[4] | Helge Holden, Nils Henrik Risebro . Follow-the-Leader models can be viewed as a numerical approximation to the Lighthill-Whitham-Richards model for traffic flow. Networks and Heterogeneous Media, 2018, 13(3): 409-421. doi: 10.3934/nhm.2018018 |
[5] | Georges Bastin, B. Haut, Jean-Michel Coron, Brigitte d'Andréa-Novel . Lyapunov stability analysis of networks of scalar conservation laws. Networks and Heterogeneous Media, 2007, 2(4): 751-759. doi: 10.3934/nhm.2007.2.751 |
[6] | Alberto Bressan, Khai T. Nguyen . Conservation law models for traffic flow on a network of roads. Networks and Heterogeneous Media, 2015, 10(2): 255-293. doi: 10.3934/nhm.2015.10.255 |
[7] | Maya Briani, Emiliano Cristiani . An easy-to-use algorithm for simulating traffic flow on networks: Theoretical study. Networks and Heterogeneous Media, 2014, 9(3): 519-552. doi: 10.3934/nhm.2014.9.519 |
[8] | Michael Herty, Niklas Kolbe, Siegfried Müller . Central schemes for networked scalar conservation laws. Networks and Heterogeneous Media, 2023, 18(1): 310-340. doi: 10.3934/nhm.2023012 |
[9] | Frederike Kissling, Christian Rohde . The computation of nonclassical shock waves with a heterogeneous multiscale method. Networks and Heterogeneous Media, 2010, 5(3): 661-674. doi: 10.3934/nhm.2010.5.661 |
[10] | Qing Li, Steinar Evje . Learning the nonlinear flux function of a hidden scalar conservation law from data. Networks and Heterogeneous Media, 2023, 18(1): 48-79. doi: 10.3934/nhm.2023003 |
In dynamical systems on networks, Kirchhoff's first law describes the local conservation of a quantity across edges. Predominantly, Kirchhoff's first law has been conceived as a phenomenological law of continuum physics. We establish its algebraic form as a property that is inherited from fundamental axioms of a network's geometry, instead of a law observed in physical nature. To this end, we extend calculus to networks, modeled as abstract metric spaces, and derive Kirchhoff's first law for hyperbolic conservation laws. In particular, our results show that hyperbolic conservation laws on networks can be stated without explicit Kirchhoff-type boundary conditions.
In the dynamical systems community, networks are largely studied as families of Euclidean edges E:=([0,we])e∈E, with we∈R>0 and E being an index set. The network's structure is often only incorporated by defining a family of dynamical systems Φe:[0,T]×E→S,e∈E, where S models some state-space, T∈R>0, and the systems are coupled by their boundary conditions. Networks typically do not appear explicitly as mathematical objects and are not subject to further study.
In this work, however, we construct networks as metric spaces (N,d), where d is a notion of path distance. We show that sufficiently regular networks (N,d) are Polish metric spaces—a convenient setting for deriving basic measure-theoretic notions underpinning abstract analysis. Although Euclidean calculus does not immediately apply to such networks, we develop, with minimal assumptions, a metric calculus for networks that preserves the properties of the known Euclidean setting.
Using such network calculus we introduce a weak notion of hyperbolic conservation laws, for which our main theorem, Theorem 4.2, states equivalence with an edge-wise system (Φe)e∈E if all Φe are a solution to a hyperbolic conservation law on a Euclidean edge and Kirchhoff's algebraic boundary conditions hold for each t∈[0,T]. The theorem applies, in particular, if a solution to a hyperbolic conservation law admits a quasi-linear form (see Remark 8). Theorem 4.2 therefore implies that an arguably universal calculus for networks exists for which hyperbolic conservation laws can be stated without explicit Kirchhoff-type boundary conditions. Concerning the uniqueness of quasi-linear hyperbolic conservation laws, an Entropy condition is required edge-wise and additional conditions need to be satisfied at the boundaries. A global analysis perspective could help here as well and one could expect to formulate an Entropy condition on the entire network without a specific emphasis on the nodes and boundary conditions.
The results of our work thereby contribute to unifying the study of global objects relevant to dynamical systems, such as time-limiting phenomena or approximations, with the study of physically-plausible transport models in a mostly Euclidean setting. Furthermore, our work derives from the most basic axioms of mathematical analysis, a widely observed and practically essential phenomenon of mathematical physics.
Motivation and related work Networks are essential models in many applications such as information technology, chemistry, power systems, transportation, neuroscience, and social sciences. In light of such broad applicability, a general theory of (dynamical systems on) networks may
● derive system properties from basic axioms, in particular, when they cannot be validated precisely by the experimental method of physics, such as in traffic system modeling,
● provide a setting for describing abstract concepts, such as Kirchhoff's law, and
● make existing mathematical results applicable, e.g., limit theorems from probability or the theory of Lp-spaces.
Rather recently, the study of networks was, therefore, extended by a global—as opposed to edgewise-Euclidean—perspective based on modeling networks as abstract metrizable spaces [12,23,25]. While such network spaces, often also called quantum graphs, have been studied greatly before [1,22,27,31,32,33], novel global phenomena may now be studied from the abstract perspective.
Transport on networks, modeled by coupled Euclidean hyperbolic conservation laws, has been of particular interest to the mathematical, science, and engineering communities [2,3,4,5,9,10,11,13,14,15,16,17,18,19,24,28,29]. Thus, we seek to derive a coupling of hyperbolic conservation laws as a consequence of the network's geometry and an associated abstract calculus characterized by universal properties. With such derivation, Kirchhoff's first law for hyperbolic conservation laws is provable and replaces a set of phenomenological model axioms.
Outline In Section 2, we introduce a particularly accessible setting for constructing networks based on graph data and equivalence relations, discuss the (un)suitability of graphs for the purpose of network modeling, rule out pathological examples of networks to develop a notion of regularity, and motivate the notion of path distance.
We continue constructing a Lebesgue-like measure for networks, developing differentiability, derivatives, and integration by parts in Section 3.
We close with motivating a notion of a weak hyperbolic conservation law for networks in Section 4, which lets us derive our main theorem, Theorem 4.2, an analog of Kirchhoff's law for these abstract dynamical systems.
Our initial goal of this work is to introduce a rigorous notion of a network. Networks ought to be the central objects on which the rest of the theory will build. A guiding perspective is taken in classical mechanics, built into many of the mathematical terms used in this section. The decisive impact of this perspective will be demonstrated first.
Due to the widespread use of graphs in discrete mathematical modeling of related real-world phenomena, it is of interest to discuss their relevance to the purpose of network modeling. Classically, only when modeling pairwise relations of elements of a set V called vertices, one resorts to graphs (V,E), with E:K→V×V, where E(k) is called an edge, K is an index set and k∈K. This representation includes multiple edges with the same vertices. Possibly, one assigns a weight W:K→R>0 to each edge to describe additional structure.
Definition 2.1 (Positively weighted graph). A positively weighted graph is a tuple (V,E,W), where V is a set, E:K→V×V is a function from an index set K to the product of V with itself, and W:K→R>0 is a function from K to the positive reals. It is required that E(K)1∪E(K)2=V, * that is, all vertices are connected to an edge.
*The subscript denotes the projection on the respective element of the tuple.
As a network model and for the purpose of modeling in this work, however, graphs will not be sufficient. In mathematical terms, one can find an indication of the unsuitability of graphs in Lemma 2.1.
Lemma 2.1. Every metric, path-connected, finite vertex space has only one vertex.
Proof. A finite metric space (V,d) has the discrete topology as for all x∈V the ball with a radius smaller than
miny∈V∖{x}d(x,y) |
contains only x. If V carries the discrete topology all subsets of V are clopen. If V is additionally path-connected, it is also connected, and thus the only clopen sets are ∅ and V. This implies V={x}.
Lemma 2.1 characterizes the incompatibility of graphs and the continuity of real numbers embedded into much of the mathematical language. This is relevant to practical problems, as continuity is the defining structure in continuum mechanics and is thus an inherent property in modeling network phenomena. In applications, the following, slightly informally stated properties are expected to be fulfilled.
A network ought to be path−connected,admit a compatible notion of distance,be complete,and have reasonable integration. | (2.1) |
Nonetheless, it turns out that one can construct networks that fulfill Property 2.1 from graphs but not as graphs—as will be shown by Definition 2.2 and Theorem 2.1. That is, positively weighted graphs non-uniquely encode the information of a specific network among all networks. Yet additional information is supplied by intervals of real numbers to obtain a general network setting with Property 2.1.
A quite elementary way to construct a network with such properties is as a set of equivalence classes of elements of intervals in R with a notion of path distance. While there are multiple ways to construct such spaces mathematically, even such that do not rely on graphs, the Definition 2.2 networks will be constructed as a set and equipped with a distance function. Some pathological examples of networks will then be ruled out in Definition 2.3, and Property 2.1 proven in Theorem 2.1. This presentation may be relatively accessible for interdisciplinary study, as it requires only a small amount of abstract language. Two examples of street networks are visualized in Figure 1.
A topological construction of networks, on the other hand, is given in [26]. Although the topological setting is arguably elegant, the presented quotient topology in general agrees with the path distance topology only for (in the work termed "combinatorically") locally finite networks (see "middle" of Figure 3). As a benefit of such a setting, it becomes clear that networks are topologically characterized by the universal property of the quotient topology. In this work, it will be shown that the assumption of a countable number of edges can be recovered from the regularity assumptions of Definition 2.3.
To define networks in a concise way, the following tools are introduced. Given a family (I(wj))j∈J of intervals I(wj):=[0,wj]⊆R with wj>0 for all j∈J and J being a non-empty index set. The disjoint union of these intervals is defined as
Un((I(wj))j∈J):={(x,j)∣x∈I(wj),j∈J}, |
and an extended metric on Un((I(wj))j∈J) is defined as
d′:Un((I(wj))j∈J)×Un((I(wj))j∈J)⟶[0,∞],((x,j),(x′,j′))⟼{|x−x′|,if j=j′∞,otherwise. |
Definition 2.2 (Network). Given a positively weighted graph (V,E,W), one can define an equivalence relation∼on the disjoint union Un((I(W(k)))k∈K) by
(x,k)∼(x′,k′):⟺{x=x′andE(k)=E(k′)orx=x′=0andE(k)1=E(k′)1orx=0andx′=W(k′)andE(k)1=E(k′)2orx=W(k)andx′=0andE(k)2=E(k′)1orx=W(k)andx′=W(k′)andE(k)2=E(k′)2. |
A network associated with (V,E,W) is defined as the set N of equivalence classes
N:={[x]∣x∈Un((I(W(k)))k∈K)}, |
where
[x]:={y∈Un((I(W(k)))k∈K)∣y∼x}. |
A network is endowed with a function
d:N×N⟶[0,∞],(x,x′)⟼inf{n∑i=1d′(pi,qi)|n∈N;(pi)i∈N≤n,(qi)i∈N≤n∈Un((I(W(k)))k∈K)x∼p1;x′∼qn;∀i∈N[1,n−1]:qi∼pi+1}. |
Remark 1 (Network). In N, the equivalence classes of the boundaries of intervals are called vertices of the network and denoted with V. The set E is called edges of the network N, where an element is a set that contains the equivalence classes of elements with the same index. The vertices of an edge e∈E are denoted by 0e and we. Note that the distance function d does not regard the orientation of the edges. The edges that contain the vertex v will be denoted with Ev.
Essentially, intervals are "glued" together, and a universal extension from distances along intervals to distances along simple paths is constructed. At this point, paths have not been defined rigorously. In fact, paths will be recovered from the distance function d—which will be shown to be a metric—in the case of regular networks.
A classic example of a network is the Wheatstone network, which originated from electrodynamics [34] and is seminal in the study of transportation systems [7].
Example 2.1 (Wheatstone network). As per Definition 2.2, one can characterize a network by a positively weighted graph (Vwh,Ewh,Wwh), where Vwh:={1,…,6}, Kwh:={1,…,7}, and Ewh:Kwh→Vwh×Vwh, and
Ewh(1):=(1,2),Ewh(2):=(2,3),Ewh(3):=(2,4),Ewh(4):=(3,4),Ewh(5):=(3,5),Ewh(6):=(4,5), andEwh(7):=(5,6). |
The weight function can be picked rather arbitrarily, but an embedding into R2 preserves vertex distances, e.g., for Wwh≡1. The network Nwh characterized by the graph (Vwh,Ewh,Wwh) is called Wheatstone network. See Figure 2 for a visualization of the Wheatstone network.
There are certain networks, depicted in Figure 3, that do not carry the structure required in Property 2.1. These networks turn out to be somewhat pathological and do not lend themselves well to the purpose of modeling real-world networks. However, the following definition describes networks that fulfill all of the requirements of Property 2.1.
Definition 2.3 (Regularity). A network N is called regular if
● it is locally finite, i.e., each vertex is included in at most a finite number of edges,
● there exists a positive lower bound of the edge lengths, and
● it is connected with respect to the topology induced by d.
Regular networks allow the description of Lebesgue integration, continuous functions, and paths. These constructs are essential to the theory and practice of modeling transport phenomena on networks. The following theorem, therefore, describes several mathematical qualities of networks that enable their study.
Theorem 2.1 (Regularity). Regular networks (N,d) are path-connected, complete, locally compact, separable metric spaces with a countable number of edges.
Remark 2 (Regularity). Regularity, therefore, implies that networks are Polish spaces. All networks considered will be regular for the remainder of this work.
Proof.
1. It is shown that d is an extended metric (may take the value ∞) on a regular network.
Identitity of indiscernibles: Let x,y∈N. If x≠y their distance d(x,y) is not smaller than the length of a subinterval of some edge. To see why that is, consider the following cases.
x is a vertex: Either y must be an inner point of one of the finite number of edges connected to x. In this case (in slight abuse of notation) d(x,y)≥min{|x−y|,ε}, where ε is a positive lower bound to the edge lengths. Otherwise, y is contained in an edge not connected to x. In this case d(x,y)≥ε.
x is not a vertex: d(x,y) is larger or equal to the shortest distance of x to the closest vertex (see previous case) or (in slight abuse of notation) d(x,y)=|x−y| if x and y are on the same edge.
If x=y, one has d(x,y)=0 as for p1∼x, q1∼y there is the extended metric distance d′(p1,q1)=0.
Symmetry: By the reversal of finite sequences and commutativity of addition, d is symmetric.
Triangle inequality: Pick sequences (pi,qi)i∈N,(p′i,q′i)i∈N and let x,y,z∈N and for γ>0 (see Definition 2.2) with
(n∑i=1d′(pi,qi))−d(x,y)<γ2and(n∑i=1d′(p′i,q′i))−d(y,z)<γ2,x∼p1,y∼qn,y∼p′1,z∼q′n,and∀i∈N<n:qi∼pi+1,q′i∼p′i+1. |
Such sequences exist by definition of d. By concatenation of the sequences, one has
d(x,z)−d(x,y)−d(y,z)<γ. |
As this holds for all γ>0, the triangle inequality follows.
2. It is shown that regular networks (N,d) are path-connected. By assumption, regular networks are connected. Regular networks are also locally path-connected by the argument following below. Generally, topological spaces that are connected and locally path-connected are path-connected. All of the above arguments are with respect to the topology induced by the extended metric d.
Let x∈N, then one can pick ε>0 such that Uε(x) contains only elements of N that share an edge with x. This is done such that Uε(x) is only a subset of one edge if x is not a vertex. Otherwise, ε should be picked as a lower bound of the edge lengths. Uε(x) is path-connected as edges are path-connected, and its restriction onto each edge is a subinterval.
3. By the following argument, one can show using the path-connectedness of N that the extended metric d takes only finite values, i.e., that d is a metric.
For x,y∈N pick a path γ:[0,1]→N with γ(0)=x and γ(1)=y. As [0,1] is compact, γ([0,1]) is also compact and in particular bounded. Therefore, d(x,y) is bounded by the same bound.
4. A regular network (N,d) has a countable number of edges.
Let ε>0 be the positive lower bound of the edge lengths and x∈V. The initial goal is to show that for all n∈N0
Aεn:={y∈V∣d(x,y)≤εn} |
is finite. This can be proved by an inductive argument. Clearly, |Aε|=0 and thus assume that Aεn is finite for some n∈N0. The vertices in Aε(n+1)∖Aεn are connected to Aεn by single edges, as per the lower bound of edge lengths ε. However, as Aεn is finite and all v∈Aεn are only connected to a finite number of edges, one has
|Aε(n+1)|≤∑v∈Aεn|Ev|<∞. |
This implies that by the countable union of finite sets, the set of vertices ⋃n∈N0Aεn is finite, and the set of edges is countable.
5. A regular network (N,d) is a complete space, as for all Cauchy sequences (xn)n∈N∈N one can pick N∈N such that (xn+N)n∈N is contained in a finite number of edges, which by restriction, is therefore included in a compact and thus complete subset.
6. A regular network (N,d) is locally compact, as it is complete and neighborhoods contain bounded closed neighborhoods.
7. Generally countable unions of separable spaces are separable. Thus, regular networks are separable as they are countable unions of edges, which are separable.
Lastly, for the following work, a precise notion of a path and its length is useful as it relates d and shortest paths.
Definition 2.4 (Path). A path in a regular network N is a continuous function p:[0,l]→N, where l>0. The length of a path is defined as
ℓ(p):=sup{n−1∑i=1d(p(ti),p(ti+1))|0=t1≤⋯≤tn=l,n∈N}. |
A path p:[0,l]→N is called parameterized by its length if ℓ(p|[0,l′])=l′ for all l′≤l.
Remark 3 (Path). Given an edge ek∈E, in slight abuse of notation ℓ(ek):=W(k) will also be written for any k∈K, where K is the index set of the graph associated with the network. For a path p:[0,l]→N, if p(0)=x and p(l)=y, it is said that x⇝y for p.
Theorem 2.2. In a regular network N, for any x,y∈N, there exists at least one shortest path x⇝y parametrized by its length d(x,y).
Proof. For x,y∈N, set ε:=d(x,y). The lower bound on edge lengths implies that there is a finite number of paths x⇝y in N∩Uε(x) that are parametrized by the length of each edge they can be restricted to. One of such that is shortest exists in this finite set of paths and must have length d(x,y) by definition of d.
One may desire a medium to localize a notion of mass in networks and describe concentrations thereof. This mass may be a homogeneous object, unlike, e.g., a set of distinguished agents. This perspective is particularly meaningful in what one may call "egalitarian systems", as those of infrastructure or platform providers. Mathematical measure theory provides a framework to study such objects and includes a notion of signed, discrete, and continuous media.
There is a unique measure on a network that assigns each edge and subinterval thereof its length. Measures of this type are essential for classical integration and the modeling of densities. As regular networks are metric spaces, one can consider their Borel σ-algebra B(N) as a set of measurable sets. An example of a discrete measure often encountered in practice in the form of data is pictured in Figure 4.
Definition 3.1 (Lebesgue measure). A measure λ is called the Lebesgue measure on a regular network (N,B(N)) if for all U⊆N that are isometrically isomorphic to a possibly degenerate interval, i.e. U≅[0,l], λ|U is the classical Lebesgue measure.
Remark 4 (Isometric isomorphism). Two metric spaces (U,dU),(V,dV) are isometrically isomorphic if there exists a bijective function f:U→V that preserves the metric distance, i.e., for all u1,u2∈U
dU(u1,u2)=dV(f(u1),f(u2)). |
One writes U≅V.
Lemma 3.1. On regular networks, a unique Lebesgue measure exists and is σ-finite.
Proof. For all elements x∈N, one finds a neighborhood Ux of x that is isometrically isomorphic to a possibly degenerate interval. Due to the compactness of single edges and the countability of the edges (see Theorem 2.1), one can find a countable set of elements B⊂N such that
⋃x∈BUx=N. |
Without loss of generality, assume that (Ux)x∈B are disjoint, as otherwise a disjoint cover can be constructed from (Ux)x∈B by subtraction and intersection of sets. Therefore, by σ-additivity every such measure λ must for all A∈B(N) satisfy the equation
λ(A)=λ(⋃x∈BA∩Ux)=∑x∈Bλ(A∩Ux)=∑x∈Bλ|Ux(A∩Ux), |
where λ|Ux is the classical Lebesgue measure on the interval Ux. The right-hand serves as the unique definition of a networks' Lebesgue measure, as for other covers of N one can show equality by refinement of the cover (Ux)x∈B. The Lebesgue measure is clearly σ-finite on regular networks as Ux includes only a vertex and has measure zero, or is included in an edge with a finite length.
The Lebesgue measure can be used to define an integral, called the Lebesgue integral (for the Lebesgue measure). The construction of this integral, a standard procedure, will not be covered. One also obtains Lebesgue densities, which will capture the notion of media defined by locally integrable functions and represent the state space of the dynamical systems studied in Section 4.
Some of the key uses of derivatives of mathematical objects are linear approximation and the representation of (physical) structures through differential equations. It is essential for this work to have a notion of differentiability for networks.
The common Euclidean theory of differentiation does not apply to networks as they are abstract metric spaces with non-oriented edges and vertices as non-trivial boundary. Therefore, it is tough to find a general notion of differentiability. The notions of differentiability differ depending on how much one likes to think of vertices as a boundary. The following definition may be the best choice since it enforces continuity constraints on values a function takes on vertices. Its offer of the existence of continuous differentiability may be essential to many use cases.
Definition 3.2 (Differentiability). A function f:N→R on a regular network is called k-(continuously) differentiable if f∘p is k-times (continuously) differentiable for all isometric paths p:[0,l]→N.
This notion of differentiability requires differentiability along all "structured" paths, yet does not require a notion of orientation, as only for derivatives one needs to pick a (local) orientation. Next to differentiability along edges, the definition turns out to impose an additional constraint of zero-derivatives on all vertices with more than two edges and differentiability across vertices with two edges (see Figure 5).
Differentiable functions are still missing a derivative. For a derivative, one has to pick an orientation for the edges. The orientation of network edges can be adapted to the orientation of a graph's edges that encodes the network, but it does not have to be. A problem that may appear is that there are vertices with two edges for which one picks different orientations. The following global object can solve this problem.
Definition 3.3 (Spatial derivative). An operator D:C1(N)→C(N) is called spatial derivative if it resembles a classical derivative, i.e., if for all open U⊆N for which there is an isometric isomorphism p:(0,l)→U it holds that for all f∈C1(N)
(Df)|U∘p=(f|U∘p)′. |
Remark 5 (Orientation of 1-d components).
1. In the setting of dynamical systems, the notation (⋅)x:=ddx:=D will also be used.
2. Each spatial derivative provides a specific assignment of 0e and we for each edge e∈E, as each edge inherits an orientation that is embedded into the definition. This fact can be characterized by monotonicity of functions: A function f∈C1(N) is said to be strictly monotonous on e if (Df)|e>0. Then the definition
0e:=arg min f|eandwe:=arg max f|e |
does not depend on the choice of f.
3. Given D, one can define the incoming and the outgoing edges of vertex v as
Einv:={e∈E∣we=v}andEoutv:={e∈E∣0e=v}. |
4. A spatial derivative can be computed by a simple reduction to the real-valued case. Indeed, for all e∈E there exists a unique path p parametrized by length 0e⇝we. As p is locally an isometric isomorphism, one has for all x∈e
Df(x)=(f∘p)′(p−1(x)). |
One obtains such a derivative by splitting the network on all vertices with more than two edges and picking an orientation for all parts. The partition's remaining elements are either isometric to intervals, or the original network is (topologically) a sphere. Piecewise derivatives are then compatible as they are zero at vertices with more than two edges, as this is required as per the differentiability of functions.
An analog of the classical product rule and the related integration by parts formula also holds for all spatial derivatives of a regular network. These properties are essential when dealing with certain dynamical systems that describe the transport of media in Section 4.
Lemma 3.2 (Product rule). A spatial derivative D:C1(N)→C(N) fulfills the product rule, i.e., for all x∈e∈E and f,g∈C1(N)
D(fg)(x)≡(fg|e)′(x)=f′|e(x)g|e(x)+f|e(x)g′|e(x)≡Df(x)g(x)+f(x)Dg(x). |
Proof. Straightforward.
Theorem 3.1 (Integration by parts). Let D:C1(N)→C(N) be a spatial derivative on a regular network N that is compact. Then, for all f,g∈C1(N)
∫NDfg+fDgdλ=∑e∈E(fg)(we)−(fg)(0e)=∑v∈V(∑e∈Einv(fg)(we)−∑e∈Eoutv(fg)(0e)), |
where 0e,we are the endpoints of edge e the order of which depends on D.
Proof. Let f,g∈C1(N) and D:C1(N)→C(N). One has
∫NDfg+fDgdλ=∑e∈E∫e(fg)′|edλ|e (Lemmas 3.1 and 3.2)=∑e∈E(fg)(we)−(fg)(0e), (fundamental theorem) |
where 0e,we are the endpoints of e, the order of which depends on D. One obtains the second identity in the theorem's statement because each edge is incoming into exactly one vertex and outgoing from exactly one vertex.
Weak solutions to differential equations can be a generalization of the notion of a solution from continuously differentiable functions to Lebesgue densities. In the case of networks, this generalization manifests as both a generalization to a non-Euclidean setting and a generalization to less regular initial data.
In Figure 6, it can be seen that even in "simple" networks, there may not exist a differentiable or even continuous solution to a conservation law. This fact and the possible generalization to less regular initial measures, such as L1loc(N), suggest a weak solution theory for networks. This setting will turn out to yield a coupled system of weak solutions for each edge and thereby recover Kirchhoff's first law from the network's topology.
The following conservation law captures the dynamical systems of Lebesgue densities that will be treated in this work.
Remark 6 (Notation). For functions ρ:[0,T]×N→R and A⊆N, the notation ρ|A:=ρ|(0,T)×A will be used. Note that in the following for functions, unless otherwise specified, (⋅)|⋅ will denote the restriction, (⋅)⋅ a partial derivative, and (⋅)⋅_ the projection onto a factor of the codomain.
For example, for a continuously differentiable function
f:R2→R2,(x,y)↦(a(x,y),b(x,y)) |
it will be written
f2_!=b,f|[0,1]2!=(x∈[0,1]2↦f(x)),andfx!=∂∂xf. |
The time integrals ∫⋅dt considered will be Lebesgue integrals. Further, A∘ will denote the interior of A⊆N.
Definition 4.1 (Conservation laws on networks). The functions
ρ∈C([0,T],L1loc(N))andν∈L∞loc([0,T]×N), |
fulfill a conservation law with a spatial derivative ddx on the regular network N, if for all ϕ∈C∞c([0,T]×N)
∬ |
It is important to mention that the above solution concept enforces the flow to be zero on vertices included in only one edge. That is, there is no in- or outflow of the network. These edges may be seen as dead-ins or dead-ends. Excluding them is a purely technical feature that offers a slightly more condensed description. Even in this setting, one can always introduce "dummy edges" and not lose generalization.
To validate the introduced solution concept, it should be made sure it is compatible with the classical setting in the sense of Theorem 4.1.
Theorem 4.1 (Classical solutions). In the setting of Definition 4.1, if has only one edge and , then fulfills a conservation law with and if and only if
Proof. Assume fulfills a conservation law with and . Let be a network with only one edge and and , then
The previous equation must hold for all ; in particular if is zero at the domain's boundary. Therefore, must be zero everywhere by the fundamental lemma of the calculus of variations [20,p. 6,Lemma 1.1.1]. If , then through arbitrary localization on the boundaries and one obtains by
and the fundamental lemma of the calculus of variations that
The opposite direction of the statement is trivial.
One of the central ideas towards simulating and optimizing network flows is their characterization as flows on edges coupled by the conservation of their boundary flow at mutual vertices. This reduces the presented abstract setting to a more classical one that turns out to lend itself well to computation. To this end, Theorem 4.2 can be understood to recover Kirchhoff's first law [21] for conservation laws of Definition 4.1. To get a sense of the meaning of boundary evaluations of -functions, the following lemma is derived in preparation of Theorem 4.2.
Lemma 4.1 (Boundary evaluations). Let
● for some ,
● ,
● for all with
(4.1) |
and such that there exists with
(4.2) |
By completeness define . Then,
Remark 7 (Right-limits for boundary evaluations). denotes uniformly continuous functions. In the following, the limit
will be heavily used in boundary evaluations of functions that fulfill the regularity assumptions of Lemma 4.1. Note that such limits may differ in network vertices based on the edge in which the limit is considered.
Proof. Let and pick such that for all
One has for all
Theorem 4.2 (Kirchhoff's first law). In the setting of Definition 4.1, if for all edges
(4.3) |
then fulfills a conservation law with and if and only if
Proof. The proof is broken down into several steps.
● Let fulfill a conservation law with and . Then, by extension of test functions from edges to the network with , it is clear that is a weak solution on edges.
● To show that the flow is conserved at vertices let and such that the support of is contained in an -ball around at all times, where is smaller than the smallest edge length, i.e.,
Note that can be contracted with a continuous bijection uniquely, such that,
For all define by
Generally, as fulfills a conservation law with and
(4.4) |
● Therefore, by the fact that
with being the number of edges connected to , and the latter factor due to possible self-edges, it follows that
(4.5) |
By
one has
(4.6) |
● One therefore has
(by Equations (4.4) to (4.6))
(linearity)
(Lemma 4.1, see below argument)
( was arbitrary, fundamental lemma c.v. [20,p. 6,Lemma 1.1.1])
To see why Lemma 4.1 is applicable for all , without loss of generalization, assume ; if , the two halves of the edge need to be treated separately. In the notation of Lemma 4.1 let
Then, by Eq (4.3) and the fact that
it remains to show that Eqs (4.1) and (4.2) hold. By Theorem 3.1 and the definition of , Eq (4.1) holds, i.e., for all
Again by definition of , Eq (4.2) holds, i.e.,
(by )
● The converse statement of this theorem holds by the following argument. Assume that is a weak solution on edges with flow conservation at all vertices. The goal is to show that for all
(4.7) |
It is equivalent to verify Eq (4.7) for test functions of type , where is a localized function, such that,
and for a vertex and (as above)
This is because can be part of a partition of unity such that all other terms are zero by the assumption of being a weak solution on open edges. Therefore, the value of
is constant in and converges as —by the exact analog derivation as above —to the difference of incoming and outgoing flow, which is zero by assumption.
Remark 8 (Generality of the solution concept). It is essential to note that Theorem 4.2 does not make a statement on the existence or uniqueness of solutions. The solution concept at hand relaxes the widely known setting of the existence of a solution for a given to the assumption of the existence of a solution tuple . In particular, there may be a representation . Therefore, the statement of Theorem 4.2 applies in greater generality, in particular, to hyperbolic conservation laws that admit a quasi-linear form. Generalizations to vector-valued conservation laws modeling multi-commodity flows as well as to balance laws are possible. An introduction to hyperbolic conservation laws can be found in [8].
We see the most promising focus of future work building on the described theory in
● Robust solutions to hyperbolic conservation laws: The fact that regular networks are Polish spaces suggests the application of existing tools from probability to meaningfully model scenarios, e.g., hyperbolic conservation laws with random initial data;
● Convergence of solutions in time: Long-time behavior/stability of physically-plausible hyperbolic conservation laws on networks could be studied for its dependence on the network geometry; and
● Localization in networks: Uncertain measurements of characteristics of hyperbolic conservation laws on networks, which are embedded into a higher-level space, could model real-world localization problems. Such models would include information about the network geometry in the localization problem.
The authors declare that they have not used Artificial Intelligence (AI) tools in the creation of this article.
The authors declare that there is no conflicts of interest.
[1] | L. Alon, R. Band, G. Berkolaiko, Universality of nodal count distribution in large metric graphs, Exp Math, (2022), 1–35. https://doi.org/10.1080/10586458.2022.2092565 |
[2] | G. Bastin, J. M. Coron, Exponential stability of networks of density-flow conservation laws under PI boundary control, IFAC Proceedings Volumes, 46 (2013), 221–226. https://doi.org/10.3182/20130925-3-FR-4043.00029 |
[3] |
G. Bastin, B. Haut, J. M. Coron, B. d'Andrea-Novel, Lyapunov stability analysis of networks of scalar conservation laws, Netw. Heterog. Media, 2 (2007), 751–759. https://doi.org/10.3934/nhm.2007.2.751 doi: 10.3934/nhm.2007.2.751
![]() |
[4] |
A. Bayen, J. Friedrich, A. Keimer, L. Pflug, T. Veeravalli, Modeling multilane traffic with moving obstacles by nonlocal balance laws, SIAM J. Appl. Dyn. Syst., 21 (2022), 1495–1538. https://doi.org/10.1137/20M1366654 doi: 10.1137/20M1366654
![]() |
[5] |
A. Bayen, A. Keimer, E. Porter, M. Spinola, Time-continuous instantaneous and past memory routing on traffic networks: A mathematical analysis on the basis of the link-delay model, SIAM J. Appl. Dyn. Syst., 18 (2019), 2143–2180. https://doi.org/10.1137/19M1258980 doi: 10.1137/19M1258980
![]() |
[6] |
G. Boeing, Osmnx: New methods for acquiring, constructing, analyzing, and visualizing complex street networks, Comput Environ Urban Syst, 65 (2017), 126–139. https://doi.org/10.1016/j.compenvurbsys.2017.05.004 doi: 10.1016/j.compenvurbsys.2017.05.004
![]() |
[7] |
D. Braess, Über ein Paradoxon aus der Verkehrsplanung, Unternehmensforschung, 12 (1968), 258–268. https://doi.org/10.1007/BF01918335 doi: 10.1007/BF01918335
![]() |
[8] | A. Bressan, Hyperbolic Systems of Conservation Laws: The One-dimensional Cauchy Problem, Oxford: Oxford University Press, 2000. |
[9] |
A. Bressan, S. Čanić, M. Garavello, M. Herty, B. Piccoli, Flows on networks: recent results and perspectives, EMS Surv Math SCI, 1 (2014), 47–111. https://doi.org/10.4171/emss/2 doi: 10.4171/emss/2
![]() |
[10] |
G. M. Coclite, M. Garavello, Vanishing viscosity for traffic on networks, SIAM J. Math. Anal., 42 (2010), 1761–1783. https://doi.org/10.1137/090771417 doi: 10.1137/090771417
![]() |
[11] |
G. M. Coclite, M. Garavello, B. Piccoli, Traffic flow on a road network, SIAM J. Math. Anal., 36 (2005), 1862–1886. https://doi.org/10.1137/S0036141004402683 doi: 10.1137/S0036141004402683
![]() |
[12] | M. Düfel, J. B. Kennedy, D. Mugnolo, M. Plümer, M. Täufer, Boundary conditions matter: On the spectrum of infinite quantum graphs, arXiv: 2207.04024, [Preprint], (2022), [cited 2023 Nov 13]. Available from: https://doi.org/10.48550/arXiv.2207.04024 |
[13] | M. Garavello, K. Han, B. Piccoli, Models for Vehicular Traffic on Networks, Springfield Missouri: American Institute of Mathematical Sciences, 2016. |
[14] |
M. Garavello, B. Piccoli, Conservation laws on complex networks, Ann. Inst. Henri Poincare (C) Anal. Non Lineaire, 26 (2009), 1925–1951. https://doi.org/10.1016/j.anihpc.2009.04.001 doi: 10.1016/j.anihpc.2009.04.001
![]() |
[15] |
F. R. Guarguaglini, R. Natalini, Global smooth solutions for a hyperbolic chemotaxis model on a network, SIAM J. Math. Anal., 47 (2015), 4652–4671. https://doi.org/10.1137/140997099 doi: 10.1137/140997099
![]() |
[16] |
M. Gugat, M. Herty, A. Klar, G. Leugering, Optimal control for traffic flow networks, J Optim Theory Appl, 126 (2005), 589–616. https://doi.org/10.1007/s10957-005-5499-z doi: 10.1007/s10957-005-5499-z
![]() |
[17] | M. Gugat, Nodal control of conservation laws on networks, In: J. Cagnol, J. P. Zolesio (Eds.) Control and boundary analysis, Boca Raton: CRC Press, 2005, 221–236. |
[18] |
M. Gugat, A. Keimer, G. Leugering, Z. Wang, Analysis of a system of nonlocal conservation laws for multi-commodity flow on networks, Netw. Heterog. Media, 10 (2015), 749–758. https://doi.org/10.3934/nhm.2015.10.749 doi: 10.3934/nhm.2015.10.749
![]() |
[19] |
H. Holden, N. H. Risebro, A mathematical model of traffic flow on a network of roads, SIAM J. Math. Anal., 26 (1995), 999–1017. https://doi.org/10.1137/S0036141093243289 doi: 10.1137/S0036141093243289
![]() |
[20] | J. Jost, X. Li-Jost, Calculus of Variations, Cambridge: Cambridge University Press, 1998. |
[21] | G. Kirchhoff, Über die Auflösung der Gleichungen, auf welche man bei der Untersuchung der linearen Vertheilung galvanischer Ströme geführt wird, Annalen der Physik, 148 (1847), 497–508. |
[22] |
T. Kottos, U. Smilansky, Periodic orbit theory and spectral statistics for quantum graphs, Ann Phys (N Y), 274 (1999), 76–124. https://doi.org/10.1006/aphy.1999.5904 doi: 10.1006/aphy.1999.5904
![]() |
[23] |
P. Kuchment, Quantum graphs: I. Some basic structures, Waves in Random Media, 14 (2003), S107. https://doi.org/10.1088/0959-7174/14/1/014 doi: 10.1088/0959-7174/14/1/014
![]() |
[24] |
N. Laurent-Brouty, A. Keimer, P. Goatin, A. M. Bayen, A macroscopic traffic flow model with finite buffers on networks: well-posedness by means of hamilton–jacobi equations, Commun Math Sci, 18 (2020), 1569–1604. https://dx.doi.org/10.4310/CMS.2020.v18.n6.a4 doi: 10.4310/CMS.2020.v18.n6.a4
![]() |
[25] | D. Mugnolo, Semigroup Methods for Evolution Equations on Networks, Cham: Springer, 2014. |
[26] | D. Mugnolo, What is actually a metric graph? arXiv: 1912.07549 [Preprint], (2022), [cited 2023 Nov 13]. Available from: https://doi.org/10.48550/arXiv.1912.07549 |
[27] |
D. Mugnolo, S. Romanelli, Dynamic and generalized wentzell node conditions for network equations, Math. Methods Appl. Sci, 30 (2007), 681–706. https://doi.org/10.1002/mma.805 doi: 10.1002/mma.805
![]() |
[28] |
L. O. Müller, P. J. Blanco, A high order approximation of hyperbolic conservation laws in networks: Application to one-dimensional blood flow, J. Comput. Phys., 300 (2015), 423–437. https://doi.org/10.1016/j.jcp.2015.07.056 doi: 10.1016/j.jcp.2015.07.056
![]() |
[29] |
M. Musch, U. S. Fjordholm, N. H. Risebro, Well-posedness theory for nonlinear scalar conservation laws on networks, Netw. Heterog. Media, 17 (2022), 101–128. https://doi.org/10.3934/nhm.2021025 doi: 10.3934/nhm.2021025
![]() |
[30] | OpenStreetMap Foundation board, OpenStreetMap contributors, Available from: https://planet.osm.org. |
[31] |
L. Pauling, The diamagnetic anisotropy of aromatic molecules, J. Chem. Phys., 4 (1936), 673–677. https://doi.org/10.1063/1.1749766 doi: 10.1063/1.1749766
![]() |
[32] |
J. R. Platt, Classification of spectra of cata‐condensed hydrocarbons, J. Chem. Phys., 17 (1949), 484–495. https://doi.org/10.1063/1.1747293 doi: 10.1063/1.1747293
![]() |
[33] |
M. Richardson, N. Balazs, On the network model of molecules and solids, Ann Phys (N Y), 73 (1972), 308–325. https://doi.org/10.1016/0003-4916(72)90285-0 doi: 10.1016/0003-4916(72)90285-0
![]() |
[34] |
C. Wheatstone, XIII. The Bakerian lecture.–An account of several new instruments and processes for determining the constants of a voltaic circuit, Philos. Trans. R. Soc. London, 133 (1843), 303–327. https://doi.org/10.1098/rstl.1843.0014 doi: 10.1098/rstl.1843.0014
![]() |