
A star shaped network with two ingoing and three outgoing edges
.Extensive research has demonstrated that cement is one of the most effective materials for improving soil properties. Researchers have investigated cement-stabilized soil techniques from various perspectives, including microstructural evolution and mechanical performance. However, studies on cement-stabilized soils in seasonal frozen regions remain limited. This study thus explored the application of cement-stabilized soil in these regions, specifically examining the effects of freeze-thaw cycles on its microstructure and shear strength through scanning electron microscopy (SEM) and direct shear tests. The findings indicate that freeze-thaw cycles induce noticeable microcracks and pores, significantly increasing particle breakage and decomposition, which leads to a loose structure and severely compromises the soil's mechanical properties. Incorporating cement generates hydration products that form cementitious bonds between soil particles, significantly enhancing structural density and overall stability. This cement stabilization effectively mitigates the damage caused by freeze-thaw cycles, enabling the soil to maintain good shear strength even after such cycles. These findings underscore the importance of cement stabilization in improving soil performance under freeze-thaw conditions, providing a theoretical basis and technical support for foundation improvement in cold regions.
Citation: Congyan Zhang, Feng Chen, Xudong Wang. Microstructural evolution and direct shear strength of cement-stabilized soil under freeze-thaw cycles[J]. AIMS Materials Science, 2025, 12(1): 28-47. doi: 10.3934/matersci.2025003
[1] | Abraham Sylla . Influence of a slow moving vehicle on traffic: Well-posedness and approximation for a mildly nonlocal model. Networks and Heterogeneous Media, 2021, 16(2): 221-256. doi: 10.3934/nhm.2021005 |
[2] | Maria Laura Delle Monache, Paola Goatin . Stability estimates for scalar conservation laws with moving flux constraints. Networks and Heterogeneous Media, 2017, 12(2): 245-258. doi: 10.3934/nhm.2017010 |
[3] | 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 |
[4] | Christophe Chalons, Paola Goatin, Nicolas Seguin . General constrained conservation laws. Application to pedestrian flow modeling. Networks and Heterogeneous Media, 2013, 8(2): 433-463. doi: 10.3934/nhm.2013.8.433 |
[5] | 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 |
[6] | Alexandre M. Bayen, Alexander Keimer, Nils Müller . A proof of Kirchhoff's first law for hyperbolic conservation laws on networks. Networks and Heterogeneous Media, 2023, 18(4): 1799-1819. doi: 10.3934/nhm.2023078 |
[7] | Jan Friedrich, Oliver Kolb, Simone Göttlich . A Godunov type scheme for a class of LWR traffic flow models with non-local flux. Networks and Heterogeneous Media, 2018, 13(4): 531-547. doi: 10.3934/nhm.2018024 |
[8] | Felisia Angela Chiarello, Giuseppe Maria Coclite . Nonlocal scalar conservation laws with discontinuous flux. Networks and Heterogeneous Media, 2023, 18(1): 380-398. doi: 10.3934/nhm.2023015 |
[9] | 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 |
[10] | 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 |
Extensive research has demonstrated that cement is one of the most effective materials for improving soil properties. Researchers have investigated cement-stabilized soil techniques from various perspectives, including microstructural evolution and mechanical performance. However, studies on cement-stabilized soils in seasonal frozen regions remain limited. This study thus explored the application of cement-stabilized soil in these regions, specifically examining the effects of freeze-thaw cycles on its microstructure and shear strength through scanning electron microscopy (SEM) and direct shear tests. The findings indicate that freeze-thaw cycles induce noticeable microcracks and pores, significantly increasing particle breakage and decomposition, which leads to a loose structure and severely compromises the soil's mechanical properties. Incorporating cement generates hydration products that form cementitious bonds between soil particles, significantly enhancing structural density and overall stability. This cement stabilization effectively mitigates the damage caused by freeze-thaw cycles, enabling the soil to maintain good shear strength even after such cycles. These findings underscore the importance of cement stabilization in improving soil performance under freeze-thaw conditions, providing a theoretical basis and technical support for foundation improvement in cold regions.
Partial differential equations (PDEs) on networks have a large number of applications, including fluid flow in pipelines, traffic flow on a network of roads, blood flow in vessels, data networks, irrigation channels and supply chains. A treatment of this wide range of applications can be found in the review articles [6,14] and the references therein. In this paper we will focus on scalar, one-dimensional conservation laws
ut+f(u)x=0 | (1.1) |
on a network. Here,
Consider a network represented by a connected and directed graph. We tag the edges of this graph with an index
ukt+fk(uk)x=0,x∈Dk, t>0uk(x,0)=ˉuk(x),x∈Dk | (1.2) |
for some spatial domain
In this paper we will be interested in uniqueness and stability for nonlinear scalar conservation laws on a network, as well as in constructing a numerical approximation and proving convergence of the numerical scheme. As opposed to many existing results, where the flux function on each edge is the same [10,18], we want to allow for a different flux function fk on each edge Dk of the network. Assuming that each flux fk is continuous and independent of the space variable, the problem can be seen as a PDE with a discontinuous flux, with the points of discontinuity sitting on the vertices of the graph. In fact, if our network would be the trivial network with only one ingoing and one outgoing edge then this would be exactly the case of a conservation law on the real line with a flux function with one discontinuity located at the vertex. Because of the connection to the theory for conservation laws with discontinuous fluxes (see e.g. [3]), we will borrow several ideas from this theory. It is well-established that nonlinear hyperbolic conservation laws develop shocks in finite time. Therefore, solutions are always understood in the weak sense. Unfortunately, weak solutions to nonlinear hyperbolic conservation laws turn out to be non-unique, and additional conditions, usually referred to as entropy conditions, are imposed to select a unique solution. If the flux function is continuous then the theory of entropy solutions is covered by Kruzkhov's theory [21]. For conservation laws with discontinuous fluxes the choice of entropy conditions is not obvious, and different physical models might lead to different entropy conditions. Although suitable entropy conditions can yield uniqueness, different entropy conditions are known to yield different solutions; see [30,19,7,3] and references therein. In [30] the author shows convergence of a finite difference scheme scheme under the assumption of a strictly concave flux function with a single maximum. In a later paper [19] a uniqueness result was shown for degenerate parabolic convection-diffusion equations, of which hyperbolic conservation laws are a subcase. In this work, the flux function had to satisfy a so-called "crossing condition". Another convergence result for an Engquist–Osher type scheme was given in [7]. The flux functions f,g were assumed to have a single maximum and to satisfy f(u)=g(u)=0 for u=0,1. The study of different entropy conditions for conservation laws with discontinuous fluxes culminated in the paper by Andreianov, Karlsen and Risebro [3]. The authors relate the question of admissibility of a solution to the properties of a set of constant solutions, a so-called germ. Inspired by the entropy theory of Andreianov, Karlsen and Risebro, we investigate so-called stationary and discrete stationary solutions for our graph problem and thus derive an entropy theory for conservation laws on networks. Although our theory is influenced by the theory in [3], we have strived to make this paper as self-contained as possible.
In [18] the authors show uniqueness and existence to the Riemann problem as well as existence of a weak solution of the Cauchy problem on a network of roads in the case that the flux functions on each edge are identical. In [9,8,1,10] the authors show well-posedness results for vanishing viscosity solutions. In [15] the authors investigate two entropy type conditions. However, in none of the existing literature can one find a general entropy condition which leads to uniqueness and stability of solutions. In the present work we aim to address this deficiency in the existing theory of conservation laws on networks.
The second important question to address is existence of a solution. Our approach will be to construct an approximation of the exact entropy solution by constructing a finite volume scheme. We will prove convergence to an entropy solution, thereby also proving existence of a solution. Convergence to the unique entropy solution of numerical schemes has been shown for conservation laws with strictly concave flux functions. This was done for schemes which are implicit on the nodes in [1,Section 3.2] and [2]. Convergence of a fully explicit scheme for the strictly concave case was shown in [29]. For a general overview over numerical methods for conservation laws on graphs see [6,Section 6].
In this article we focus on monotone fluxes – that is, each flux
This article is structured as follows: In Section 2 we define our mathematical framework. We show uniqueness of entropy solutions to our problem in Section 3. In Section 4 we define a finite difference scheme appropriate for our problem, and in Section 5 we prove that our numerical scheme converges towards the unique entropy solution. In Section 6 we show that a class of monotone flux functions fits in our general scheme. Numerical experiments for the monotone case are presented in Section 7.
While the theory outlined in Sections 2 through 4 holds for conservation laws with general flux functions, the convergence theory in Sections 5 and 7 focuses on monotone flux functions and upwind numerical fluxes.
Consider a network (or directed graph) of vertices and edges; for simplicity we will assume that the network contains a single vertex, along with
Dk={R−for k∈Iin,R+for k∈Iout. |
On each edge
ukt+fk(uk)x=0for x∈Dk, k∈I. | (2.1) |
The collection of functions u=(uk)k∈I can be thought of as a function u:Ω→R, where
Ω:=⋃k∈IDk×{k}. |
On the Borel
∫Ωudλ=∑k∈I∫Dkuk(x)dx. | (2.2) |
The set of
TV(u):=∫Ω|dudx|dλ=∑k∈I∫Dk|dukdx(x)|dx. | (2.3) |
Definition 2.1 (Weak Solution). We say that a function
∑k∈I∫∞0∫Dkukφkt+fk(uk)φkxdxdt+∑k∈I∫Dkˉuk(x)φk(x,0)dx=0 | (2.4) |
for all
Weak solutions automatically satisfy a Rankine–Hugoniot condition at the intersection:
Proposition 2.2 (Rankine–Hugoniot condition). Let
∑k∈Iinfk(uk)(0,t)=∑k∈Ioutfk(uk)(0,t)for a.e.t>0. | (2.5) |
Proof. Define
θε(x)={1ε(ε+x)if x∈[−ε,0]1ε(ε−x)if x∈[0,ε]0if |x|>ε. | (2.6) |
We define Φ(x,t):=θε(x)ψ(t) where ψ∈C∞c([0,∞)). The partial derivatives of
Φx(x,t)={1εψ(t)if x∈[−ε,0]−1εψ(t)if x∈[0,ε]0if |x|>εandΦt(x,t)=θε(x)ψ′(t). |
By a density argument, Φ qualifies as an admissible test function. Thus, we can insert Φ into the weak formulation (2.4) to get
0=∑k∈I∫∞0∫DkukΦkt+fk(uk)Φkxdxdt+∑k∈I∫Dkˉuk(x)Φk(x,0)dx=∑k∈I∫∞0∫Dkukθε(x)ψ′(t)dtdx+1ε∑k∈I∫∞0∫Dk∩(−ε,ε)sgn(k)fk(uk)ψ(t)dxdt−1ε∑k∈I∫Dk∩(−ε,ε)(ε−x)ˉuk(x,0)ψ(0)dx→−∑k∈I∫∞0sgn(k)fk(uk)ψ(t)dt |
as
Definition 2.3 (Stationary Solution). A stationary solution of (2.1) is a weak solution of (2.1) which is constant in time and is a strong solution on each edge
∑k∈Iinfk(ck)=∑k∈Ioutfk(ck). | (2.7) |
Thus, we can identify each stationary solution with a vector
Remark 2.4. Note that if we only required stationary solutions to be weak solutions on each edge
Next, we formulate conditions that will single out a unique weak solution.
Definition 2.5 (Kruzkov entropy pairs). The Kruzkov entropy pairs are the pairs of functions
The Kruzkov entropy pairs lead to a consistency condition on sets of stationary solutions:
Definition 2.6. A subset
∑k∈Iinqkck(˜ck)⩾∑k∈Ioutqkck(˜ck) | (2.8) |
for every pair
The set of stationary solutions
Definition 2.7. Let
L∞oco(G)={u∈L∞(Ω;λ) : ∃ c,d∈G s.t. ck⩽uk(x)⩽dk ∀ (x,k)∈Ω.} | (2.9) |
Example 2.8. If
L∞oco(G2)={u∈L∞(Ω;λ) : u−1≡c, u1≡−c for some c⩾0}. |
Thus,
Definition 2.9 (Entropy Solution). Let
∑k∈I∫∞0∫Dkηck(uk)φkt+qkck(uk)φkxdxdt+∑k∈I∫∞0ηck(ˉuk(x))φk(x,0)dx⩾0 | (2.10) |
for every
Audusse and Perthame [4] considered an entropy condition similar to (2.10), but in the context of spatially dependent, discontinuous flux functions.
We show first that entropy solutions are invariant in the set
Lemma 2.10. Let
Proof. Select
∑k∈I∫∞0∫Dk(ck−uk)+φkt+H(ck−uk)(fk(ck)−fk(uk))φkxdxdt+∑k∈I∫∞0(ck−ˉuk(x))+⏟=0φk(x,0)dx⩾0 |
(where
∑k∈I∫Dk(ck−uk(x,T))+dx⩽0 |
for a.e.
The above lemma enables us to show that entropy solutions have strong traces.
Lemma 2.11. Let
Proof. It follows from Lemma 2.10 that
Proposition 2.12. Let
∑k∈Iinqkck(uk)(0,t)⩾∑k∈Ioutqkck(uk)(0,t)for a.e.t>0 | (2.11) |
for every
Proof. We take a positive test function
0⩽∑k∈I∫∞0∫Dkηck(uk)θε(x)ψ′(t)dxdt−1ε∑k∈I(∫∞0∫Dk∩(−ε,ε)sgn(k)qkck(uk)ψ(t)dxdt+∫Dkηck(ˉuk(x))θε(x)ψ(t)dx)→−∑k∈I∫∞0sgn(k)qkck(uk(0,t))ψ(t)dt |
as
Corollary 2.13. If
Theorem 3.1 (Entropy Solutions are L1 stable). Let
∑k∈I‖uk(t)−vk(t)‖L1(Dk)⩽∑k∈I‖ˉuk−ˉvk‖L1(Dk) |
for every
Proof. From Lemma 2.10 it follows that
∫Dk∫∞0|uk(x,t)−vk(x,t)|φt+qkv(u)φxdtdx+∫Dk|ˉuk(x)−ˉvk(x)|φ(x,0)dx⩾0. | (3.1) |
Next, for general
μh(x):={0x∈(−∞,−2h)1h(x+2h)x∈[−2h,−h)1x∈[−h,0] |
and
Ψh(x):=1−μh(x). |
The derivative of Ψh reads
Ψ′h(x)={0x∈(−∞,−2h)−1hx∈[−2h,−h)0x∈[−h,0]. |
Define φk(x,t):=ξk(x,t)Ψh(x) for a function ξk∈C∞c(¯Dk×[0,∞)). We insert φ into equation (3.1) to get
∫Dk∫∞0|uk(x,t)−vk(x,t)|ξktΨh+qkvk(uk)ξkxΨhdtdx+∫Dk∫∞0qkvk(uk)ξkΨ′hdtdx+∫Dk|ˉuk(x)−ˉvk(x)|ξkΨhdx⩾0. |
Sending h↓0 we get
∫Dk∫∞0|uk(x,t)−vk(x,t)|ξkt+qkvk(uk)ξkxdtdx+∫Dk|ˉuk(x)−ˉvk(x)|ξkdx+limh↓0∫−h−2h∫∞0qkvk(uk)ξkΨ′hdtdx⩾0. |
Since the traces of qk(uk) and qk(vk) exist, we get
−limh↓01h∫T0∫−h−2hqkvk(uk)ξkdxdt=−∫T0qkvk−(uk)ξk(0,t)dt. |
We therefore obtain
∫Dk∫∞0|uk(x,t)−vk(x,t)|ξt+qkvk(uk)ξxdtdx+∫Dk|ˉuk(x)−ˉvk(x)|dx−∫T0qkvk(uk)ξ(0,t)dt⩾0. | (3.2) |
By an analogous argument we get
∫Dk∫∞0|uk(x,t)−vk(x,t)|ξt+qkvk(uk)ξxdtdx+∫Dk|ˉuk(x)−ˉvk(x)|dx+∫T0qkvk(uk)ξ(0,t)dt⩾0 |
for k∈Iout. Fix
αr(x)={0x∈(−∞,−r−1]x+r+1x∈(−r−1,−r)1x∈[−r,0)βκ(t)={1t∈[0,s]1κ(κ+s−t)t∈(s,s+κ)0t∈[s+κ,∞). |
Via a standard regularization argument one can check that φ(x,t)=αr(x)βκ(t) is an admissible test function. We compute the partial derivatives of φ:
φt(x,t)={0t∈[0,s]−1καr(x)t∈(s,s+κ]0t∈(s+κ,∞) |
and
φx(x,t)={0x∈(−∞,−r−1)βκ(t)x∈(−r−1,−r)0x∈(−r,0). |
We insert this into (3.2) to get
−1κ∫s+κs∫0−r−1|uk(x,t)−vk(x,t)|αr(x)dxdt+∫s+κ0∫−r−r−1qkvk(uk)βκ(t)dxdt+∫0−r−1|ˉuk(x)−ˉvk(x)|αr(x)dx−∫s+κ0qkvk(uk(0,t))βκ(t)dt⩾0. |
Letting κ→0 and r→∞, we get
‖uk(x,t)−vk(x,t)‖L1(Dk)⩽‖ˉuk(x)−ˉvk(x)‖L1(Dk)−∫s0qkvk(uk(0,t))dt. |
An analogous inequality holds for
∑k∈I‖uk(x,t)−vk(x,t)‖L1(Dk)⩽∑k∈I∫Dk|ˉuk(x)−ˉvk(x)|dx+∫s0∑k∈Isgn(k)qkvk(uk)⏟⩽0 by (2.11) and Corollary 2.13⩽∑k∈I‖ˉuk(x)−ˉvk(x)‖L1(Dk). |
In this section we construct a finite volume numerical approximation for (2.1) and prove stability and convergence properties of the method. The numerical method is rather standard for hyperbolic conservation laws, but an important feature of the method is that the vertex is discretized as a separate control volume. Although this control volume vanishes as the mesh parameter
Let
D+disc:=N,D−disc:=−N,Dkdisc:=Dsgn(k)disc,D0disc:={0}. |
For
1In numerical experiments, the timestep
Cki=Dk∩(xi−1/2,xi+1/2). |
We define the mesh size at the vertex by
uk,ni≈1Δx∫Ckiuk(x,tn)dxfor i∈Dkdisc,un0≈1Δx0∑k∈I∫Ck0uk(x,tn)dx. |
Fix some i∈Dkdisc, let
uk,n+1i−uk,niΔt+Fk,ni+1/2−Fk,ni−1/2Δx=0 | (4.1a) |
where
Fk,ni+1/2≈1Δt∫tn+1tnfk(uk(xi+1/2,t))dt. |
For the special cell
un+10−un0Δt+1Δx0(∑k∈IoutFk,n1/2−∑k∈IinFk,n−1/2)=0. | (4.1b) |
(This is opposed to the explicit method of Towers [29] where the vertex is modelled as having zero width for any Δx>0.) We will use the notational convention that
Given a numerically computed solution
uΔt(x,k,t)=uk,nifor x∈Cki, t∈[tn,tn+1). | (4.2) |
We remark that the integral of
∫ΩuΔt(⋅,t)dλ=∑k∈I∑i∈Dkdiscuk,niΔx+un0Δx0 | (4.3) |
for any
TV(uΔt(⋅,t))=∑k∈Iin∑i∈Dkdisc|uk,ni+1−uk,ni|+∑k∈Iout∑i∈Dkdisc|uk,ni−uk,ni−1|=∑k∈Iin∑i∈Dkdisc|uk,ni−uk,ni−1|+∑k∈Iout∑i∈Dkdisc|uk,ni+1−uk,ni|+∑k∈Iin|un0−uk,n−1|+∑k∈Iout|un0−uk,n1|. | (4.4) |
Note also that a numerical method of the form (4.1) is conservative in the sense that the total mass
∫ΩuΔt(⋅,tn+1)dλ=∑k∈I∑i∈Dkdiscuk,n+1iΔx+un+10Δx0=∑k∈I∑i∈Dkdiscuk,niΔx−Δt(Fk,ni+1/2−Fk,ni−1/2)+un0Δx0−Δt(∑k∈IoutFk1/2−∑k∈IinFk−1/2)=∑k∈I∑i∈Dkdiscuk,niΔx+un0Δx0=∫ΩuΔt(⋅,tn)dλ. |
As a shorthand for the scheme (4.1) we define the functions
Gk(uki−1,uki,uki+1):=uki−ΔtΔx(Fk(uki,uki+1)−Fk(uki−1,uki)) | (4.5a) |
for
G0(u−Nin−1,…,u−1−1,u0,u11,…,uNout1):=u0−ΔtΔx0(∑k∈IoutFk(u0,uk1)−∑k∈IinFk(uk−1,u0)), | (4.5b) |
enabling us to write (4.1) in the update form
uk,n+1i=Gk(uk,ni−1,uk,ni,uk,ni+1)for i∈Dkdisc, k∈Iun+10=G0(u−Nin,n−1,…,u−1,n−1,un0,u1,n1,…,uNout,n1). | (4.6) |
As a shorthand for (4.6), we will sometimes use the notation
uk,n+1i=Gk(uni−1,uni,uni+1)for i∈Dkdisc, k∈I0, | (4.6') |
where
Definition 4.1 (Monotone scheme). The difference scheme (4.6') is monotone if
un⩽vn⇒un+1⩽vn+1, |
where
We state a straightforward CFL-type condition which ensures monotonicity of the numerical scheme.
Proposition 4.2. Consider a consistent finite volume method (4.1), where
Δtmaxk,u,v|∂Fk∂u(u,v)|⩽Δx/2,Δtmaxk,u,v|∂Fk∂v(u,v)|⩽Δx/2. | (4.7) |
Proof. We can calculate the derivatives to the update functions to get
∂Gk∂uki−1=ΔtΔx∂Fki−1/2∂uki−1,∂Gk∂uki+1=−ΔtΔx∂Fki+1/2∂uki+1∂Gk∂uki=1−ΔtΔx(∂Fki+1/2∂uki−∂Fki−1/2∂uki), |
for each
∂G0∂uk−1=ΔtΔx0∂Fk−1/2∂uk−1for k∈Iin,∂G0∂uk1=−ΔtΔx0∂Fk1/2∂uk1for k∈Iout,∂G0∂u0=1−ΔtΔx0(∑k∈Iout∂Fk1/2∂un0−∑k∈Iin∂Fk−1/2∂un0) |
on the vertex. We would like these derivatives to be non-negative. The monotonicity of
∂Gk∂uki=ΔtΔx(ΔxΔt−|∂Fki+1/2∂uki|−|∂Fki−1/2∂uki|)⩾0 |
(by (4.7)) and
∂G0∂u0=ΔtΔx0(Δx0Δt−∑k∈Iout|∂Fk1/2∂un0|−∑k∈Iin|∂Fk−1/2∂un0|) |
(using Δx0=NΔx/2)
⩾ΔtΔx0(NΔx2Δt−Noutmaxk,u,v|∂Fk∂u(u,v)|−Ninmaxk,u,v|∂Fk∂v(u,v)|)⩾0 |
by (4.7).
Remark 4.3. As opposed to the explicit method that Towers proposes in [29], where the CFL condition gets more restrictive as the number of roads grows, we don't face any issues with the time step with the allowable time step with a high number of roads.
In the same way that stationary solutions are essential for the well-posedness of entropy solutions (cf. Section 3), they are essential to the stability and convergence of numerical methods on networks. Asserting that a numerical solution is both constant in time and on each edge yields the following definition.
Definition 4.4 (Discrete Stationary Solution). Consider a consistent, conservative numerical method (4.1). A discrete stationary solution for (4.1) is a vector
cdisc:=(c−Nin,…,cNout)∈RN+1 |
satisfying the Rankine–Hugoniot condition
∑k∈Iinfk(ck)=∑k∈Ioutfk(ck) | (4.8) |
as well as the conditions
Fk(ck,c0)=fk(ck)for k∈Iin, | (4.9a) |
Fk(c0,ck)=fk(ck)for k∈Iout. | (4.9b) |
In the remainder, sets of discrete stationary solutions will be denoted with a superscript,
Remark 4.5. Note that our definition of a discrete stationary solution is analogous to [1,Definition 2.1]. As opposed to our definition, the authors of [1] only include values on the edges. The value c0, which is called p in [1], is excluded from the vectors of stationary solutions there.
Notation 4.6. We will sometimes index a discrete stationary solution as
ci={(c−Nin,…,c−1)i<0c0i=0(c1,…,cNout)i>0 | (4.10a) |
for
cki={cki≠0c0i=0. | (4.10b) |
Using the notation (4.6'), it is readily checked that discrete stationary solutions are precisely those that are constant on each edge and satisfy
ci=Gk(ci−1,ci,ci+1)∀ i∈Dkdisc, k∈I0. |
Remark 4.7. The conditions (4.9) say that the numerical fluxes at the vertex reduce to the upwind flux on the in edges and the downwind flux on the out edges. This can be interpreted as information only flowing into the vertex, not out of it. This is consistent with the interpretation of the vertex as a stationary shock.
Remark 4.8. Discrete stationary solutions c=(c−Nin,…,cNout) fulfil a discrete version of the Rankine–Hugoniot type condition (2.7),
∑k∈IinFk(ck,c0)=∑k∈IoutFk(c0,ck). |
Lemma 4.9. Consider a consistent, conservative numerical scheme (4.1). Let c=(ck)k∈I be a stationary solution for (1.1) and let
c0∈⋂k∈Iin(Hk)−1({fk(ck)})⋂⋂k∈Iout(Jk)−1({fk(ck)}) |
where
Hk(c):=Fk(ck,c)fork∈Iin,Jk(c):=Fk(c,ck)fork∈Iout. |
Proof. We can rewrite conditions (4.9a) and (4.9b) as
(4.9b)⇔Hk(c0)=fk(ck)⇔c0∈(Hk)−1({fk(ck)}) |
for k∈Iin, and
(4.9a)⇔Jk(c0)=fk(ck)⇔c0∈(Jk)−1({fk(ck)}) |
for
We set out to prove an L∞ bound, L1 contractiveness and Lipschitz continuity in time for solutions computed with a general consistent, conservative, monotone finite volume method on a network. Our starting point will be a class of discrete stationary solutions
uk,0i=1Δx∫Ckiˉuk(x)dx,u00=c0. | (4.11) |
(The value
Lemma 4.10. Consider monotone numerical flux functions
Proof. Define
Ik(ck):={(Fk(ck,⋅))−1({fk(ck)})for k∈Iin(Fk(⋅,ck))−1({fk(ck)})for k∈Iout. |
Since all Fk are monotone, each Ik(ck) is a connected interval which contains ck, and moreover, Lemma 4.9 says that c0∈⋂k∈IIk(ck). This implies that
˜c0:=min(⋂k∈I[[c0,ck]]) |
exists and satisfies
˜d0:=max(⋂k∈I[[d0,dk]]), |
which satisfies
Proposition 4.11. Consider a consistent, conservative, monotone finite volume method (4.1), (4.11) with a set of discrete stationary states
Proof. Pick discrete stationary states
uk,n+1i=Gk(uni−1,uni,uni+1)⩾Gk(ci−1,ci,ci+1)=cki |
for all
Definition 4.12 (L1 contractive method). A numerical method (4.6') is L1 contractive if
‖uΔt(⋅,t)−vΔt(⋅,t)‖L1(Ω;λ)⩽‖ˉu−ˉv‖L1(Ω;λ) |
for all t⩾0, where
We state the well known Crandall–Tartar lemma which we will use in the following proof. Here and below, we use the notation
Theorem 4.13. (Crandall–Tartar: [11,Proposition 1]). Let (Ω,λ) be a measure space. Let C⊂L1(Ω;λ) have the property that f,g∈C implies f∨g∈C. Let V:C→L1(Ω;λ) satisfy ∫ΩV(f)dλ=∫Ωfdλ for f∈C. Then the following three properties of V are equivalent:
(a) f,g∈C and f⩽g a.e. implies V(f)⩽V(g) a.e.,
(b) ∫Ω(V(f)−V(g))+⩽∫Ω(f−g)+ for f,g∈C,
(c) ∫Ω|V(f)−V(g)|⩽∫Ω|f−g| for f,g∈C.
We can now prove
Theorem 4.14. Every conservative, consistent monotone method (4.1), (4.11) is L1-contractive.
Proof. Let
CΔx={u∈L1∩L∞(Ω;λ) : u(x)=∑k∈I0∑i∈Dkdiscuki1Cki for uki∈R}. |
We define the operator V:CΔx→CΔx mapping a numerical solution to the next time step,
V(u):=∑k∈I∑i∈Dkdisc1Cki(uki−ΔtΔx(Fk(uki,uki+1)−Fk(uki−1,uki)))+∑k∈I1Ck0(u00−ΔtΔx0(∑k∈IoutFk(u0,uk1)−∑k∈IinFk(uk−1,u0))). |
By the definition (2.2) of the measure
From L^1 -contractivity we get continuity in time as a corollary:
Corollary 4.15. Consider a consistent, conservative and monotone method (4.1). Let
\begin{align*} \big\| {\mathbf{u}}_ {\Delta t}(t^{n+1}) - {\mathbf{u}}_ {\Delta t}(t^n) \big\|_{L^1(\Omega;\lambda)} & \leqslant \big\| {\mathbf{u}}_ {\Delta t}(t^1) - {\mathbf{u}}_ {\Delta t}(t^0) \big\|_{L^1(\Omega;\lambda)} \\ & \leqslant {\Delta t} \big(C{\rm TV}( {\mathbf{u}}^0) + \bar M\big), \end{align*} |
where the constants
Proof. We compute
\begin{align*} \bigl\|& {\mathbf{u}}_ {\Delta t}(t^{n+1}) - {\mathbf{u}}_ {\Delta t}(t^n) \bigr\|_{L^1(\Omega;\lambda)} \\ & = \big\|V( {\mathbf{u}}_ {\Delta t}(t^{n})) - V( {\mathbf{u}}_ {\Delta t}(t^{n-1})) \big\|_{L^1(\Omega;\lambda)} \end{align*} |
(using Theorem 4.13(c))
\begin{align*} & \leqslant \big\| {\mathbf{u}}_ {\Delta t}(t^{n}) - {\mathbf{u}}_ {\Delta t}(t^{n-1}) \big\|_{L^1(\Omega;\lambda)} \leqslant \dots \leqslant \big\| {\mathbf{u}}_ {\Delta t}(t^1) - {\mathbf{u}}_ {\Delta t}(t^0) \big\|_{L^1(\Omega;\lambda)} \\ & = {\Delta x} \sum\limits_{k\in {\mathscr{I}}} \sum\limits_{i \in D_ \mathrm{disc}^k} | u_i^{k, 1} - u_i^0| + {\Delta x}_0|u_0^{k, 1} - u_0^0| \\ & = {\Delta t} \sum\limits_{k\in {\mathscr{I}}} \sum\limits_{i \in D_ \mathrm{disc}^k} \big|F_ {i+{1/2}}^{k, 0} - F_ {i-{1/2}}^{k, 0}\big| + {\Delta t} \biggl|\sum\limits_{k\in {{\mathscr{I}}_{\mathrm{out}}}} F_ {1/2}^{k, 0} - \sum\limits_{k\in {{\mathscr{I}}_{\mathrm{in}}}} F_{- {1/2}}^{k, 0}\biggr| \\ & = {\Delta x} \lambda \sum\limits_{k\in {\mathscr{I}}} \sum\limits_{i\in D_ \mathrm{disc}^k}\big|u_i^{k, 0} - u_{i-1}^{k, 0}\big| \\ &\quad+ {\Delta t} \biggl|\!\begin{aligned}[t]&\sum\limits_{k\in {{\mathscr{I}}_{\mathrm{out}}}} F_ {1/2}^{k, 0} - F^{k, 0} \big(u_0^0, u_0^0\big) - \sum\limits_{k\in {{\mathscr{I}}_{\mathrm{in}}}} F_{- {1/2}}^{k, 0} - F^{k, 0} (u_0^0, u_0^0) \\ &+\overbrace{\sum\limits_{k\in {{\mathscr{I}}_{\mathrm{out}}}} f^k \big(u_0^0\big)}^{ = : f_ {\mathrm{out}}(u_0^0)} - \overbrace{\sum\limits_{k\in {{\mathscr{I}}_{\mathrm{in}}}} f^k\big(u_0^0\big)}^{ = : f_ {\mathrm{in}}(u_0^0)}\biggr|\end{aligned} \end{align*} |
\begin{align*} & \leqslant {\Delta t} \sum\limits_{k\in {\mathscr{I}}} \sum\limits_{i\in D_ \mathrm{disc}^k} L^k \bigl(\big|u_i^{k, 0} - u_{i-1}^{k, 0}\big| + \big|u_{i+1}^{k, 0} - u_i^{k, 0}\big| \bigr) \\ &\quad + {\Delta t} \underbrace{\Biggl(\sum\limits_{k\in {{\mathscr{I}}_{\mathrm{out}}}} L^k \big|u_1^{k, 0} - u_0^0\big| + \sum\limits_{k\in {{\mathscr{I}}_{\mathrm{in}}}} L^k \big|u_0^0 - u_{-1}^{k, 0}\big| + \overbrace{\big|f_ {\mathrm{out}}\big(u_0^0\big) - f_ {\mathrm{in}}\big(u_0^0\big)\big|}^{ \leqslant M}\Biggr)}_{ = :\bar M} \\ & \leqslant {\Delta t} (C {\rm TV}( {\mathbf{u}}^0) + \bar M ), \end{align*} |
where we collect all constants into the global constant C . We can bound \big|f_ {\mathrm{out}} \big(u_0^0\big) - f_ {\mathrm{in}} \big(u_0^0\big)\big| \leqslant \bar M with a constant \bar M \in \mathbb{R} since f_ {\mathrm{in}}, f_ {\mathrm{out}} are continuous and u_0^0 \in L^\infty .
We are now in place to prove convergence in the case where the flux functions
F^k(u, v) = \begin{cases} f^k(u) & \text{if $f^k$ is increasing, } \\ f^k(v) & \text{if $f^k$ is decreasing.} \end{cases} |
We shall show that the set of discrete approximations is compact in
Theorem 5.1 (Lax–Wendroff theorem). Fix T > 0 . Assume that each flux function f^k is locally Lipschitz continuous and strictly monotone. Let
Remark 5.2. The existence of a non-trivial mutually consistent germ \mathscr{G}_ \mathrm{disc}^0 for monotone flux functions will be shown in 6.
Proof. We write
Let
\begin{align*} Q_ {i+{1/2}}^{k, n} = F^k \bigl( u_i^{k, n} \vee c^k_i, u_{i+1}^{k, n} \vee c^k_i \bigr) - F^k \bigl( u_i^{k, n} \wedge c^k, u_{i+1}^{k, n} \wedge c^k \bigr) \end{align*} |
for
\begin{align*} Q_{- {1/2}}^n = \sum\limits_{k\in {{\mathscr{I}}_{\mathrm{in}}}} Q_{- {1/2}}^{k, n}, \qquad Q_{ {1/2}}^n = \sum\limits_{k\in {{\mathscr{I}}_{\mathrm{out}}}} Q_{ {1/2}}^{k, n} \end{align*} |
(cf. Notation 4.6 for the definition of
\begin{align*} & G^k \bigl( u_{i-1}^{k, n} \vee c^k, u_i^{k, n} \vee c^k , u_{i + 1}^{k, n} \vee c^k \bigr) - G^k \bigl( u_{i-1}^{k, n} \wedge c^k, u_i^{k, n} \wedge c^k , u_{i + 1}^{k, n} \wedge c^k \bigr)\\ & \qquad = \bigl|u_i^{k, n} - c^k\bigr| - \frac{ {\Delta t}}{ {\Delta x}} \bigl( Q_ {i+{1/2}}^{k, n} - Q_ {i-{1/2}}^{k, n} \bigr), \end{align*} |
for
\begin{equation} \begin{split} \big|u_i^{k, n+1}-c^k\big| & = u_i^{k, n+1}\vee c^k - u_i^{k, n+1}\wedge c^k \\ & = G^k \bigl(u_{i-1}^{k, n}, u_i^{k, n}, u_{i + 1}^{k, n}\bigr)\vee c^k - G^k \bigl(u_{i-1}^{k, n}, u_i^{k, n}, u_{i + 1}^{k, n}\bigr)\wedge c^k \\ & \leqslant G^k \bigl( u_{i-1}^{k, n} \vee c^k, u_i^{k, n} \vee c^k , u_{i + 1}^{k, n} \vee c^k \bigr) \\ & \quad- G^k \bigl( u_{i-1}^{k, n} \wedge c^k, u_i^{k, n} \wedge c^k , u_{i + 1}^{k, n} \wedge c^k \bigr) \\ & = \bigl|u_i^{k, n} - c^k\bigr| - \frac{ {\Delta t}}{ {\Delta x}} \bigl( Q_ {i+{1/2}}^{k, n} - Q_ {i-{1/2}}^{k, n} \bigr). \end{split} \end{equation} | (5.1) |
Similarly, we find that
\begin{equation} \bigl|u_0^{n+1} - c^0\bigr| - \bigl|u_0^n - c^0\bigr| + \frac{ {\Delta t}}{ {\Delta x}_0} \bigl( Q_ {1/2}^n - Q_{- {1/2}}^n \bigr) \leqslant 0 \end{equation} | (5.2) |
We choose T = M {\Delta t} for a natural number M , multiply the above
\begin{align*} 0& \geqslant\sum\limits_{n = 0}^{M}\sum\limits_{k\in {\mathscr{I}}} \sum\limits_{i \in D_ \mathrm{disc}^k} \varphi_i^{k, n}\biggl( \bigl( \big|u_i^{k, n+1} - c^k\big| - \big|u_i^{k, n} - c^k\big| \bigr) + \frac{ {\Delta t}}{ {\Delta x}} \bigl( Q_ {i+{1/2}}^{k, n} - Q_ {i-{1/2}}^{k, n} \bigr)\biggr) \\ &\quad +\sum\limits_{n = 0}^{M} \varphi_0^n\biggl( \frac{N}{2} \bigl( \big|u_0^{n+1} - c^0\big| - \big|u_0^n - c^0\big| \bigr) + \frac{ {\Delta t}}{ {\Delta x}} \biggl( \sum\limits_{k\in {{\mathscr{I}}_{\mathrm{out}}}} Q_{ {1/2}}^{k, n} - \sum\limits_{k\in {{\mathscr{I}}_{\mathrm{in}}}} Q_{- {1/2}}^{k, n} \biggr) \biggr), \end{align*} |
where
\begin{align*} 0& \geqslant-\sum\limits_{n = 1}^{M} \sum\limits_{k\in {\mathscr{I}}} \sum\limits_{i \in D_ \mathrm{disc}^k} \big|u_i^{k, n} - c^k\big| \bigl( \varphi_i^{k, n} - \varphi_i^{k, n-1} \bigr) - \sum\limits_{k\in {\mathscr{I}}} \sum\limits_{i \in D_ \mathrm{disc}^k} \big|u_i^{k, 0} - c^k\big| \varphi_i^{k, 0} \\ &\quad -\frac{ {\Delta t}}{ {\Delta x}} \sum\limits_{n = 0}^{M} \biggl(\begin{aligned}[t]& \sum\limits_{k\in {{\mathscr{I}}_{\mathrm{in}}}} \sum\limits_{i \in D_ \mathrm{disc}^k} Q_ {i-{1/2}}^{k, n} \bigl( \varphi_i^{k, n} - \varphi_{i-1}^{k, n} \bigr) \\ &+ \sum\limits_{k\in {{\mathscr{I}}_{\mathrm{out}}}} \sum\limits_{i \in D_ \mathrm{disc}^k} Q_ {i+{1/2}}^{k, n} \bigl( \varphi_{i+1}^{k, n} - \varphi_i^{k, n} \bigr) \\ &+ \sum\limits_{k\in {{\mathscr{I}}_{\mathrm{in}}}} \varphi_{-1}^{k, n} Q_{- {1/2}}^{k, n} - \sum\limits_{k\in {{\mathscr{I}}_{\mathrm{out}}}} \varphi_1^{k, n} Q_{ {1/2}}^{k, n} \biggr)\end{aligned} \\ &\quad-\frac{N}{2} \big|u_0^n - c^0\big| \varphi_0^0 - \frac{N}{2} \sum\limits_{n = 1}^{M} \big|u_0^n - c^0\big| \bigl( \varphi_0^n - \varphi_0^{n-1} \bigr) \\ &\quad+ \frac{ {\Delta t}}{ {\Delta x}} \sum\limits_{n = 0}^{M} \biggl(\sum\limits_{k\in {{\mathscr{I}}_{\mathrm{out}}}} Q_{ {1/2}}^{k, n} \varphi_0^n - \sum\limits_{k\in {{\mathscr{I}}_{\mathrm{in}}}} Q_{- {1/2}}^{k, n} \varphi_0^n \biggr). \end{align*} |
After shifting the
\begin{align*} 0& \geqslant-\underbrace{ {\Delta t} {\Delta x} \sum\limits_{n = 1}^{M} \sum\limits_{k\in {\mathscr{I}}} \sum\limits_{i \in D_ \mathrm{disc}^k} \big|u_i^{k, n} - c^k\big| \biggl( \frac{ \varphi_i^{k, n} - \varphi_i^{k, n-1}}{ {\Delta t}} \biggr)}_{ = \, A_1} \end{align*} |
\begin{align*} &\quad- \underbrace{ {\Delta x} \sum\limits_{k\in {\mathscr{I}}} \sum\limits_{i \in D_ \mathrm{disc}^k} \big|u_i^{k, 0} - c^k\big| \varphi_i^{k, 0}}_{ = \, A_2} \\ &\quad- \underbrace{ {\Delta t} {\Delta x} \sum\limits_{n = 0}^{M} \biggl(\begin{aligned}[t]&\sum\limits_{k\in {{\mathscr{I}}_{\mathrm{in}}}} \sum\limits_{i \in D_ \mathrm{disc}^k} Q_ {i+{1/2}}^{k, n} \biggl( \frac{ \varphi_{i+1}^{k, n} - \varphi_{i}^{k, n}}{ {\Delta x}} \biggr) \\ &+ \sum\limits_{k\in {{\mathscr{I}}_{\mathrm{out}}}} \sum\limits_{i \in D_ \mathrm{disc}^k} Q_ {i-{1/2}}^{k, n} \biggl( \frac{ \varphi_{i}^{k, n} - \varphi_{i-1}^{k, n}}{ {\Delta x}} \biggr)\biggr)\end{aligned}}_{ = \, A_3} \\ &\quad- \underbrace{ {\Delta x} \frac{N}{2} \big|u_0^{0} - c^0\big| \varphi_0^0 - {\Delta t} {\Delta x} \frac{N}{2} \sum\limits_{n = 1}^{M} \big|u_0^n - c^0\big| \biggl( \frac{ \varphi_0^n - \varphi_0^{n-1}}{ {\Delta t}} \biggr)}_{ = \, A_4}. \end{align*} |
Taking limits we get for {\Delta t}, {\Delta x} \to 0
\begin{align*} A_1 \to \sum\limits_{k\in {\mathscr{I}}} \int_0^{\infty} \int_{D^k} \big|u^k - c^k\big| \varphi_t^k \, dx \, dt, \end{align*} |
and for {\Delta x} \to 0
\begin{align*} A_2 \to \sum\limits_{k\in {\mathscr{I}}} \big|u_0^k - c^k\big|(x) \varphi^k(x, 0) \, dx, \qquad A_4 \to 0. \end{align*} |
Thus, we are left with
\begin{align*} A_3 & = {\Delta t} {\Delta x} \sum\limits_{n = 0}^{M} \biggl(\begin{aligned}[t]&\sum\limits_{k\in {{\mathscr{I}}_{\mathrm{in}}}} \sum\limits_{i \in D_ \mathrm{disc}^k} q_{c^k}^k \bigl( u_i^{k, n} \bigr) \biggl( \frac{ \varphi_{i+1}^{k, n} - \varphi_{i}^{k, n}}{ {\Delta x}} \biggr) \\ &+ \sum\limits_{k\in {{\mathscr{I}}_{\mathrm{out}}}} \sum\limits_{i \in D_ \mathrm{disc}^k} q_{c^k}^k \bigl( u_{i}^{k, n} \bigr) \biggl( \frac{ \varphi_{i}^{k, n} - \varphi_{i-1}^{k, n}}{ {\Delta x}} \biggr)\biggr) \end{aligned}\\ &\to \sum\limits_{k\in {\mathscr{I}}} \int_0^T \int_{D^k} q_{c^k}^k \bigl( u^k(x, t) \bigr) \varphi_x (x, t) \, dx \, dt \end{align*} |
as {\Delta t}, {\Delta x} \to 0 , due to the a.e. pointwise convergence of
Now we have everything in place to proof a compactness theorem.
Theorem 5.3 (Compactness and Convergence to a Weak Solution). Fix T > 0 . Assume that each flux function f^k is locally Lipschitz continuous and strictly monotone. Let
Proof. We first show convergence of the sequence of functions {\mathbf{g}}_ {\Delta t}\colon\Omega\times[0, T]\to \mathbb{R} ,
{\mathbf{g}}_ {\Delta t}(x, k, t): = f^k(u^k_ {\Delta t}(x, t)). |
The sequence
\begin{align*} \int_{\Omega} \big| {\mathbf{g}}_ {\Delta t}(t^{n+1})- {\mathbf{g}}_ {\Delta t}(t^n)\big|\, d\lambda & \leqslant C_f\int_{\Omega}\big| {\mathbf{u}}_ {\Delta t}(t^{n+1})- {\mathbf{u}}_ {\Delta t}(t^n)\big|\, d\lambda \\ & \leqslant C_f(C{\rm TV}(\bar {\mathbf{u}})+\bar M) {\Delta t}, \end{align*} |
by Corollary 4.15. We can bound the total variation of
\begin{align*} {\rm TV}( {\mathbf{g}}_ {\Delta t}(\cdot, t)) & = \sum\limits_{k\in {\mathscr{I}}}\sum\limits_{i \in D_ \mathrm{disc}^k} \big|f^k( u_i^{k, n}) - f^k( u_{i-1}^{k, n})\big| \\ &\quad + \sum\limits_{k\in {{\mathscr{I}}_{\mathrm{in}}}}\big|f^k(u_0^n)-f^k(u_{-1}^{k, n})\big| \\ & \leqslant \sum\limits_{k\in {\mathscr{I}}}\sum\limits_{i \in D_ \mathrm{disc}^k} \big|f^k( u_i^{k, n}) - f^k( u_{i-1}^{k, n})\big| + N\| {\mathbf{g}}_ {\Delta t}\|_{\infty} \\ & = \sum\limits_{k\in {\mathscr{I}}} \sum\limits_{i \in D_ \mathrm{disc}^k} \big|F_ {i+{1/2}}^{k, n} - F_ {i-{1/2}}^{k, n}\big| + N\| {\mathbf{g}}_ {\Delta t}\|_{\infty} \\ & = \frac{ {\Delta x}}{ {\Delta t}}\sum\limits_{k\in {\mathscr{I}}} \sum\limits_{i \in D_ \mathrm{disc}^k}\big|u_i^{k, n+1}-u_i^{k, n}\big| + N\| {\mathbf{g}}_ {\Delta t}\|_{\infty} \\ & \leqslant {\Delta x}(C{\rm TV}(\bar {\mathbf{u}}) + \bar M) + N\| {\mathbf{g}}_ {\Delta t}\|_{\infty}. \end{align*} |
Applying Ascoli's compactness theorem together with Helly's theorem, we get the existence of a subsequence
{\mathbf{u}}_ {\Delta t}(x, k, t) = ( f^k )^{-1}\big( {\mathbf{g}}_ {\Delta t}(x, k, t)\big), |
and hence, also
So far we have shown that if a sufficiently large class of stationary and discrete stationary solutions exists, then our equations on the network are well posed and the finite volume numerical approximations converge to the entropy solution. In this section we show that such classes exist in the case where either all fluxes f^k are strictly increasing or all are strictly decreasing. We also remark on the more general case.
We henceforth assume that all fluxes are increasing; one can attain analogous results for decreasing fluxes following the same arguments. In the following we want to investigate the sets of discrete stationary solutions implied by the upwind method.
We define
\begin{equation*} f_ {\mathrm{in}}(u) : = \sum\limits_{k\in {{\mathscr{I}}_{\mathrm{in}}}} f^k(u), \qquad f_ {\mathrm{out}}(u) : = \sum\limits_{k\in {{\mathscr{I}}_{\mathrm{out}}}} f^k(u) \qquad \text{for } u\in \mathbb{R}. \end{equation*} |
It is clear that f_ {\mathrm{in}}, f_ {\mathrm{out}} are monotone by the monotonicity of their summand components. In particular, the two functions are invertible.
For the upwind method the conditions (4.9a) and (4.9b) become
\begin{align} f^k ( c^k ) & = f^k ( c^k ) \qquad \text{for } k\in {{\mathscr{I}}_{\mathrm{in}}}, \end{align} | (6.1a) |
\begin{align} f^k ( c^0 ) & = f^k ( c^k ) \qquad \text{for }k\in {{\mathscr{I}}_{\mathrm{out}}}. \end{align} | (6.1b) |
This is equivalent to
\begin{align} c^k & = c^k \qquad \text{for } k\in {{\mathscr{I}}_{\mathrm{in}}}, \end{align} | (6.2a) |
\begin{align} c^0 & = c^k \qquad \text{for }k\in {{\mathscr{I}}_{\mathrm{out}}}, \end{align} | (6.2b) |
due to the invertibility of the flux functions f^k . It is obvious as well, that for two different discrete stationary solutions {\mathbf{c}}, {\mathbf{d}} satisfying c^k \leqslant d^k for k \in {\mathscr{I}} , we also have
\mathscr{G}_ \mathrm{disc}^0 : = \bigl\{\text{all discrete stationary states for the upwind method}\bigr\} |
and we let
\mathscr{G}_ \mathrm{disc} : = \bigl\{\big(c^{-N_ {\mathrm{in}}}, \dots, c^{-1}, c^1, \dots, c^{N_ {\mathrm{out}}}\big) \mid {\mathbf{c}}\in \mathscr{G}_ \mathrm{disc}^0)\bigr\}. |
Although it might be too difficult to find a full characterization of the set
I_ {\mathrm{in}} : = f_ {\mathrm{in}}^{-1} (R_ {\mathrm{in}}\cap R_ {\mathrm{out}}), \qquad I_ {\mathrm{out}} : = f_ {\mathrm{out}}^{-1} (R_ {\mathrm{in}}\cap R_ {\mathrm{out}}) |
where
R_ {\mathrm{in}} : = f_ {\mathrm{in}}( \mathbb{R}), \qquad R_ {\mathrm{out}} : = f_ {\mathrm{out}}( \mathbb{R}). |
By the continuity of
Theorem 6.1. We have
\mathscr{L}: = \Bigl\{ {\mathbf{u}}\in L^\infty(\Omega;\lambda) \mid u^k(x) \in I_ {\mathrm{in}}\ \forall\ k\in {{\mathscr{I}}_{\mathrm{in}}}, \ u^k(x)\in I_ {\mathrm{out}}\ \forall\ k\in {{\mathscr{I}}_{\mathrm{out}}} \Bigr\}. |
In particular, if
Proof. Let
\underline{c}_ {\mathrm{in}}: = \inf\limits_{\substack{x\in D^k\\ k\in {{\mathscr{I}}_{\mathrm{in}}}}}u^k(x) \in I_ {\mathrm{in}}, \qquad \overline{c}_ {\mathrm{in}}: = \sup\limits_{\substack{x\in D^k\\ k\in {{\mathscr{I}}_{\mathrm{in}}}}}u^k(x) \in I_ {\mathrm{in}} |
and likewise for
In a similar way one finds a stationary solution
\underline{d}_ {\mathrm{in}} \leqslant u^k(x) \leqslant\overline{d}_ {\mathrm{in}} \;\forall\ k\in {{\mathscr{I}}_{\mathrm{in}}} \qquad \text{and} \qquad \underline{d}_ {\mathrm{out}} \leqslant u^k(x) \leqslant\overline{d}_ {\mathrm{out}} \;\forall\ k\in {{\mathscr{I}}_{\mathrm{out}}} |
we conclude that
Proposition 6.2. Consider a conservation law on a network with strictly increasing fluxes f^k . Let \mathscr{G}_ \mathrm{disc}^0 denote the set of all discrete stationary solutions for the upwind method. Then the set
\mathscr{G}_ \mathrm{disc} : = \Big\lbrace \big(c^{- {N_{\mathrm{in}}}}, \dots, c^{-1}, c^1, \dots, c^{ {N_{\mathrm{out}}}}\big) \mid {\mathbf{c}} \in \mathscr{G}^0_ \mathrm{disc} \Big\rbrace |
is a mutually consistent and maximal set of stationary solutions.
Proof. Every
To prove mutual consistency of \mathscr{G}_ \mathrm{disc} we plug a discrete stationary solution {\mathbf{d}}\in \mathscr{G}_ \mathrm{disc}^0 into (5.1) to get for n \in \mathbb{N} ,
Q_{- {1/2}}^{k, n} \geqslant Q_{- {3/2}}^{k, n} \;\text{for }k \in {\mathscr{I}}_ {\mathrm{in}} \quad\text{and}\quad Q_{ {3/2}}^{k, n} \geqslant Q_ {1/2}^{k, n} \;\text{for } k \in {\mathscr{I}}_ {\mathrm{out}}. |
Since we are using the upwind scheme, this reduces to
q_{c^k}^k (d^k) \geqslant q_{c^0}^k (d^0) \text{ for } k \in {\mathscr{I}}_ {\mathrm{in}} \qquad\text{and}\qquad q_{c^0}^k (d^0) \geqslant q^k_{c^k} (d^k) \text{ for } k \in {\mathscr{I}}_ {\mathrm{out}}. |
In the same manner, plugging d^0 into (5.2) gives us
\sum\limits_{k\in {{\mathscr{I}}_{\mathrm{in}}}} Q_{- {1/2}}^{k, n} \geqslant \sum\limits_{k\in {{\mathscr{I}}_{\mathrm{out}}}} Q_ {1/2}^{k, n}. |
Combining these two observations, we get
\sum\limits_{k\in {{\mathscr{I}}_{\mathrm{in}}}} q_{c^k}^k (d^k) \geqslant \sum\limits_{k\in {{\mathscr{I}}_{\mathrm{in}}}} q_{c^0}^k (d^0) \geqslant \sum\limits_{k\in {{\mathscr{I}}_{\mathrm{out}}}} q_{c^0}^k (d^0) \geqslant \sum\limits_{k\in {{\mathscr{I}}_{\mathrm{out}}}} q_{c^k}^k (d^k). |
As
If for some vector {\mathbf{d}} , the set \mathscr{G}_ \mathrm{disc} \cup \lbrace {\mathbf{d}} \rbrace is mutually consistent, then
\sum\limits_{k\in {\mathscr{I}}_ {\mathrm{in}}} q^k_{c^k}\big(d^k\big) \geqslant \sum\limits_{k\in {\mathscr{I}}_ {\mathrm{out}}} q^k_{c^k}\big(d^k\big). |
We choose c^k = d^k\ \forall\ k \in {\mathscr{I}}_ {\mathrm{in}} and c^0 = (f_ {\mathrm{out}})^{-1}(\sum_{k\in {\mathscr{I}}_ {\mathrm{in}}} f^k(c^k)) . Since all f^k are monotonically increasing, the entropy flux reduces to q^k_{c^k}(d^k) = \big|f^k(c^k) - f^k(d^k)\big| and thus,
0 \geqslant \sum\limits_{k\in {\mathscr{I}}_ {\mathrm{out}}} \big|f^k(c^0) - f^k(d^k)\big|, |
which implies d^k = c^0 for k \in {\mathscr{I}}_ {\mathrm{out}} , and thus,
Although the framework presented in this manuscript is only applied to monotone flux functions, we remark here on the generalization of our results to more general choices of
● compactness of the sequence of approximations (here achieved via a TV bound on the (upwind) numerical flux);
● the existence of a maximal set of stationary states, and the consistency of the approximations with respect to that set.
A TV bound on the numerical fluxes can be achieved in a more general setting, but that does not easily translate to compactness of the approximation itself. This can be achieved by a detour via the Temple functional [28]. The derivation of a maximal set of stationary states requires a careful design of the numerical method. We address both of these issues in the upcoming paper [13], where we prove convergence of an Engquist–Osher-type finite volume method for more general flux functions.
We show numerical experiments for some example cases including results for linear and nonlinear as well as convex and concave fluxes. In all experiments we use a CFL number of
Example 7.3 | Example 7.5 | Example 7.4 | Example 7.1 | Example 7.2 | ||||||
Grid level | EOC | EOC | EOC | EOC | EOC | |||||
3 | 0.10877 | - | 0.11630 | - | 0.14459 | - | 0.07087 | - | 0.09904 | - |
4 | 0.05496 | 0.98 | 0.07136 | 0.70 | 0.08016 | 0.85 | 0.0546 | 0.38 | 0.04913 | 1.01 |
5 | 0.03649 | 0.59 | 0.04372 | 0.71 | 0.04651 | 0.79 | 0.03117 | 0.81 | 0.02844 | 0.79 |
6 | 0.02629 | 0.47 | 0.02255 | 0.96 | 0.02711 | 0.78 | 0.01903 | 0.71 | 0.01627 | 0.81 |
7 | 0.01830 | 0.52 | 0.01360 | 0.73 | 0.01495 | 0.86 | 0.01115 | 0.77 | 0.00919 | 0.82 |
8 | 0.01255 | 0.54 | 0.00653 | 1.06 | 0.00925 | 0.69 | 0.00644 | 0.79 | 0.00527 | 0.80 |
9 | 0.00883 | 0.51 | 0.00325 | 1.01 | 0.00480 | 0.95 | 0.00330 | 0.96 | 0.00268 | 0.98 |
10 | 0.00625 | 0.50 | 0.00160 | 1.02 | 0.00295 | 0.70 | 0.00173 | 0.93 | 0.00150 | 0.84 |
11 | 0.00442 | 0.50 | 0.00086 | 0.90 | 0.00152 | 0.96 | 0.00085 | 1.03 | 0.00084 | 0.84 |
12 | 0.00312 | 0.50 | 0.00040 | 1.10 | 0.00081 | 0.91 | 0.00042 | 1.02 | 0.00047 | 0.84 |
Example 7.1 (Burgers' equation with roundabout). In this example we include a roundabout – an edge whose endpoints meet at the same vertex, as shown in Figure 2. This case was not included in the theory but is interesting because it is analogous to a periodic boundary condition. We also include an ingoing edge and two outgoing edges, amounting to a total of two ingoing and three outgoing edges. As initial data we choose constants on the roundabout and the outgoing edges and two different constants on the independent ingoing edge. After a while the shock in the initial data on the independent ingoing edge will reach the edge and create new Riemann problems. We choose the initial data
\begin{align*} \bar u^{-1}(x) = \begin{cases} 2 & \text{if } 0 \leqslant x < 0.5, \\ \sqrt{2} & \text{if } 0.5 \leqslant x, \end{cases}\qquad \bar u^{-2} = \bar u^1 = \bar u^2 = \bar u^3 \equiv 1. \end{align*} |
We take all edges to have length
\begin{equation*} u^{-1}(x, t) = \begin{cases} 2 & \text{if } 0 \leqslant x < \frac{1}{2-\sqrt{2}} t \\ \sqrt{2} & \text{if } \frac{1}{2-\sqrt{2}} t \leqslant x \end{cases} \end{equation*} |
which will hit the vertex at t^* = 1-\frac{1}{\sqrt{2}} . To compute the solution after
\begin{align*} u^k(x, t^*) = \begin{cases} \sqrt{5/3} & \text{if } x = 0, \\ 1 & \text{if } 0 < x \end{cases} \end{align*} |
for k = 1, 2, 3 , which results in a travelling shock wave with speed s = \frac{1}{\sqrt{5/3} - 1} . At time t^{**} : = \sqrt{5/3} - \frac{1}{\sqrt{2}} the travelling shock wave which originated on the roundabout edge hits the vertex once again, resulting in a new set of Riemann problems on the outgoing edge. This process will continue in a periodic fashion.
We compute up to time
Example 7.2. We construct an example where we take the flux function from the traffic flow example in [18], f(u) = 4u(1-u) , but allow for different fluxes on different edges, f^k(u) = \alpha^k f \big(\frac{u}{\alpha^k} \big) for
Solving the conditions (4.8), (4.9) for
c^0 = \frac{3 \pm \sqrt{9 - 4 \bigl(\frac{1}{\alpha_1} + \frac{1}{\alpha_2} + \frac{1}{\alpha_3}\bigr) \big( u^{-1} - \frac{1}{\alpha_{-1}} ( u^{-1} )^2 + u^{-2} - \frac{1}{\alpha_{-2}} ( u^{-2} )^2 \big)}}{2\big(\frac{1}{\alpha_1} + \frac{1}{\alpha_2} + \frac{1}{\alpha_3}\big)}. |
For the incoming edges to have a monotonically increasing flux we impose u_{-i} \leqslant \frac{1}{2} \alpha_{-i}, for i = 1, 2 and for outgoing edges c^0 \leqslant \frac{1}{2} \min \lbrace \alpha_1, \alpha_2, \alpha_3 \rbrace . We choose \alpha_{-1} = \alpha_{-2} = 1 , \alpha_1 = \alpha_2 = 4 and \alpha_3 = 2 with initial data
\begin{align*} \bar u^{-1} = \bar u^{-2} \equiv 0.5, \qquad \bar u^1 \equiv 0, \qquad \bar u^2 \equiv \frac{1}{2} \big(3-\sqrt{7}\big), \qquad \bar u^3 \equiv 1. \end{align*} |
This gives us c^0 = \frac{1}{2}(3 - \sqrt{7}) . On the outer boundary we choose zero Neumann boundary conditions. For u^3 we will get a shock
u^3(x, t) = \begin{cases} \frac{1}{2}(3 - \sqrt{7}) & \text{if } x < st, \\ 0 & \text{if } x \geqslant st, \end{cases} |
with speed s = \frac{(3-\sqrt{7})(2-\frac{3-\sqrt{7}}{4})-3}{\frac{3-\sqrt{7}}{2}-1}, and a rarefaction wave for u^1 of the form
u^1(x, t) = \begin{cases} \frac{3-\sqrt{7}}{2} & \text{if }x < 2 (\sqrt{7}-1)t, \\ 1-\frac{x}{4t} & \text{if } 2(\sqrt{7}-1)t \leqslant x < 4t, \\ 0 & \text{if } 4t \leqslant x. \end{cases} |
On edge
We compute up to time
In addition to the examples described above we show errors and experimental order of convergence (EOC) for several additional examples in Table 1.
Example 7.3 (EOC: Linear advection). We consider a linear advection equation with two ingoing edges and three outgoing edges as in Figure 1 with initial data
\begin{align*} \bar u^{-1}(x, t) = \begin{cases} 2 & 0 \leqslant x < 0.8, \\ 1 & 0.8 \leqslant x, \end{cases}\qquad \bar u^{-2} \equiv 1, \qquad \bar u^1 = \bar u^2 = \bar u^3 \equiv \frac{2}{3}, \end{align*} |
and Dirichlet boundary conditions adapted to the edge values. We initialize the vertex node by
Example 7.4 (EOC: Burgers' equation with elementary waves). We choose \bar u^{-1} = \bar u^{-2} \equiv 1 as initial data on the ingoing roads and \bar u^1 \equiv 0 , \bar u^2 \equiv\sqrt{2/3} and \bar u^3 \equiv 2 on the outgoing edges of a star shaped graph as in Figure 1. The conditions on the numerical flux imply then 3(c^0)^2 = 2 \Leftrightarrow c^0 = \sqrt{3/2} . Thus, we get the following Riemann problems on the outgoing roads:
\begin{align*} \bar u^1(x) = \begin{cases} \sqrt{2/3} & x = 0, \\ 0 & x > 0, \end{cases}\qquad \bar u^2 = \begin{cases} \sqrt{2/3} & x = 0, \\ \sqrt{2/3} & x > 0, \end{cases}\qquad \bar u^3 = \begin{cases} \sqrt{2/3} & x = 0, \\ 2 & x > 0, \end{cases} \end{align*} |
with zero Neumann boundary conditions at the outer edges. The solution to these problems are a shock, a constant solution and a rarefaction wave, respectively. We compute up to time
Example 7.5 (EOC: Burgers' equation with travelling shock). We consider a Burgers-type equation with two ingoing edges and three outgoing edges as in Figure 1 with initial data
\begin{align*} \bar u^{-1}(x) = \begin{cases} 2 & 0 \leqslant x < 0.8, \\ 1 & 0.8 \leqslant x, \end{cases}\qquad \bar u^{-2} \equiv 1, \qquad \bar u^1 = \bar u^2 = \bar u^3 \equiv \sqrt{\frac{2}{3}}, \end{align*} |
with Dirichlet boundary conditions of the same value as the associated edge. On the vertex node the initial condition is chosen as
Convergence order estimates for finite volume methods for nonlinear scalar conservation laws are due to Kuznetsov [22] for the continuous flux case and due to Badwaik, Ruf [5] for the case of monotone fluxes with points of discontinuity. In both of those cases the analytically proven convergence rate is at least \sqrt{ {\Delta x}} . Our numerical experiments indicate the same lower bound on the convergence rate for our numerical methods on graphs. Considering the fact that f_ {\mathrm{in}} and f_ {\mathrm{out}} from Section 6 are monotone it might be possible to generalize the result of Badwaik and Ruf to networks.
In conclusion we have defined a framework for the analysis and numerical approximation of conservation laws on networks. We extended the concepts well known from the conventional case such as weak solution, entropy solution and monotone methods to make sense on a directed graph. We defined a reasonable entropy condition under which we have shown stability and uniqueness of an analytic solution. Existence is shown by convergence of a conservative, consistent, monotone difference scheme. In an upcoming work [13] we want to address convergence of a numerical method where the fluxes
We would like to thank the referee for the valuable comments, helping to improve the quality of this work.
[1] | Carter JP, Liu MD (2005) Review of the structured cam clay model, In: Jerry A. Yamamuro PE, Victor NK, Soil Constitutive Models: Evaluation, Selection, and Calibration, Reston: American Society of Civil Engineers Publishing, 99–132. https://doi.org/10.1061/40771(169)5 |
[2] |
Guo Y, Yu X (2016) Characterizing the surface charge of clay minerals with Atomic Force Microscope (AFM). AIMS Mater Sci 4: 582–593. https://doi.org/10.3934/matersci.2017.3.582 doi: 10.3934/matersci.2017.3.582
![]() |
[3] |
Lyu H, Gu J, Li W, et al. (2020) Analysis of compressibility and mechanical behavior of red clay considering structural strength. Arab J Geosci 13: 411. https://doi.org/10.1007/s12517-020-05352-4 doi: 10.1007/s12517-020-05352-4
![]() |
[4] |
Zhang Y, Yang G, Chen W, et al. (2022) Relation between microstructures and macroscopic mechanical properties of earthen-site soils. Materials 15: 6124. https://doi.org/10.3390/ma15176124 doi: 10.3390/ma15176124
![]() |
[5] |
Hu Q, Song W, Hu J (2023) Study of the mechanical properties and water stability of microbially cured, coir-fiber-reinforced clay soil. Sustainability 15: 13261. https://doi.org/10.3390/su151713261 doi: 10.3390/su151713261
![]() |
[6] |
Ge S, Zang J, Wang Y, et al. (2021) Combined stabilization/solidification and electroosmosis treatments for dredged marine silt. Mar Georesour Geotechnol 39: 1157–1166. https://doi.org/10.1080/1064119X.2020.1817205 doi: 10.1080/1064119X.2020.1817205
![]() |
[7] |
Bahadori H, Hasheminezhad A, Taghizadeh F (2019) Experimental study on marl soil stabilization using natural pozzolans. J Mater Civ Eng 31: 04018363. https://doi.org/10.1061/(ASCE)MT.1943-5533.0002577 doi: 10.1061/(ASCE)MT.1943-5533.0002577
![]() |
[8] |
Peethamparan S, Olek J, Diamond S (2009) Mechanism of stabilization of Na-montmorillonite clay with cement kiln dust. Cem Concr Res 39: 580–589. https://doi.org/10.1016/j.cemconres.2009.03.013 doi: 10.1016/j.cemconres.2009.03.013
![]() |
[9] |
Bahadori H, Hasheminezhad A, Mohamadi Asl S (2022) Stabilisation of urmia lake peat using natural and artificial pozzolans. Proc Inst Civ Eng Ground Improv 175: 104–113. https://doi.org/10.1680/jgrim.19.00024 doi: 10.1680/jgrim.19.00024
![]() |
[10] |
Song Y, Geng Y, Dong S, et al. (2023) Study on mechanical properties and microstructure of basalt fiber-modified red clay. Sustainability 15: 4411. https://doi.org/10.3390/su15054411 doi: 10.3390/su15054411
![]() |
[11] |
Petry TM, Little DN (2002) Review of stabilization of clays and expansive soils in pavements and lightly loaded structures-history, practice, and future. J Mater Civ Eng 14: 447–460. https://doi.org/10.1061/(ASCE)0899-1561(2002)14:6(447) doi: 10.1061/(ASCE)0899-1561(2002)14:6(447)
![]() |
[12] |
Medvey B, Dobszay G (2020) Durability of stabilized earthen constructions: A review. Geotech Geol Eng 38: 2403–2425. https://doi.org/10.1007/s10706-020-01208-6 doi: 10.1007/s10706-020-01208-6
![]() |
[13] |
Anburuvel A (2023) The engineering behind soil stabilization with additives: a state-of-the-art review. Geotech Geol Eng 42: 1–42. https://doi.org/10.1007/s10706-023-02554-x doi: 10.1007/s10706-023-02554-x
![]() |
[14] |
Lemaire K, Deneele D, Bonnet S, et al. (2013) Effects of lime and cement treatment on the physiochemical, microstructural and mechanical characteristics of a plastic silt. Eng Geol 166: 255–261. http://dx.doi.org/10.1016/j.enggeo.2013.09.012 doi: 10.1016/j.enggeo.2013.09.012
![]() |
[15] |
Miraki H, Shariatmadari N, Ghadir P, et al. (2022) Clayey soil stabilization using alkali-activated volcanic ash and slag. J Rock Mech Geotech Eng 14: 576–591. https://doi.org/10.1016/j.jrmge.2021.08.012 doi: 10.1016/j.jrmge.2021.08.012
![]() |
[16] |
Fonseca A, Cruz R, Consoli N (2009) Strength properties of sandy soil-cement admixtures. Geotech Geol Eng 27: 681–686. http://dx.doi.org/10.1007/s10706-009-9267-y doi: 10.1007/s10706-009-9267-y
![]() |
[17] |
Mitchell, J, El Jack S (1966) The fabric of soil-cement and its formation. Clays Clay Miner 14: 297–305. https://doi.org/10.1346/CCMN.1966.0140126 doi: 10.1346/CCMN.1966.0140126
![]() |
[18] | Islam M, Hashim R (2009) Bearing capacity of stabilized tropical peat by deep mixing method. Aust J Basic Appl Sci 3: 682–688. http://www.ajbasweb.com/old/ajbas/2009/682-688 |
[19] |
Horpibulsuk S, Rachan R, Suddeepong A (2011) Assessment of strength development in blended cement admixed bangkok clay. Constr Build Mater 25: 1521–1531. https://doi.org/10.1016/j.conbuildmat.2010.08.006 doi: 10.1016/j.conbuildmat.2010.08.006
![]() |
[20] |
Du C, Zhang J, Yang G, et al. (2021) The influence of organic matter on the strength development of cement-stabilized marine soft clay. Mar Georesour Geotech 39: 983–993. https://doi.org/10.1080/1064119X.2020.1792593 doi: 10.1080/1064119X.2020.1792593
![]() |
[21] |
Horpibulsuk S, Miura N, Bergado DT (2004) Undrained shear behavior of cement admixed clay at high water content. J Geotech Geoenviron Eng ASCE 130: 1096–1105. https://doi.org/10.1061/(ASCE)1090-0241(2004)130:10(1096) doi: 10.1061/(ASCE)1090-0241(2004)130:10(1096)
![]() |
[22] | Horpibulsuk S, Miura N (2001) A new approach for studying behavior of cement stabilized clays. Proceedings of the 15th international conference on soil mechanics and geotechnical engineering (ISSMGE), Istanbul, Turkey, 1759–1762. Available from: https://www.issmge.org/publications/online-library. |
[23] |
Miura N, Horpibulsuk S, Nagaraj TS (2001) Engineering behavior of cement stabilized clay at high water content. Soils Found 41: 33–45. https://doi.org/10.3208/sandf.41.5_33 doi: 10.3208/sandf.41.5_33
![]() |
[24] |
Horpibulsuk S, Miura N, Nagaraj TS (2005) Clay-water/cement ratio identity of cement admixed soft clay. J Geotech Geoenviron Eng ASCE 131: 187–192. https://doi.org/10.1061/(ASCE)1090-0241(2005)131:2(187) doi: 10.1061/(ASCE)1090-0241(2005)131:2(187)
![]() |
[25] |
Zhou L, Shen X, Bai Z, et al. (2009) Experimental study on the effect of external admixtures on the modification of cemented soils. Ind Constr 39: 74–78 (in Chinese). https://doi.org/10.13204/j.gyjz2009.07.022 doi: 10.13204/j.gyjz2009.07.022
![]() |
[26] |
Wang J, Ding G, Pan L, et al. (2010) Study on mechanical properties and constitutive modeling of cemented soils in static triaxial tests. Rock Soil Mech 31: 1407–1412 (in Chinese). https://doi.org/10.16285/j.rsm2010.05.043 doi: 10.16285/j.rsm2010.05.043
![]() |
[27] |
Shihata S, Baghdadi Z (2001) Simplified method to assess freeze-thaw durability of soil cement. J Mater Civ Eng 13: 243–247. https://doi.org/10.1061/(ASCE)0899-1561(2001)13:4(243) doi: 10.1061/(ASCE)0899-1561(2001)13:4(243)
![]() |
[28] |
Bouassida M, Porbaha A (2004) Ultimate bearing capacity of soft clays reinforced by a group of columns-application to a deep mixing technique. Soils Found 44: 91–101. https://doi.org/10.3208/sandf.44.3_91 doi: 10.3208/sandf.44.3_91
![]() |
[29] |
Horpibulsuk S, Liu D, Liyanapathirana S, et al. (2010) Behaviour of cemented clay simulated via the theoretical framework of the structured cam clay model. Comput Geotech 37: 1–9. https://doi.org/10.1016/j.compgeo.2009.06.007 doi: 10.1016/j.compgeo.2009.06.007
![]() |
[30] |
Jongpradist P, Youwai S, Jaturapitakkul C (2010) Effective void ratio for assessing the mechanical properties of cement-clay admixtures at high water content. J Geotech Geoenviron Eng 137: 621–627. https://doi.org/10.1061/(ASCE)GT.1943-5606.0000462 doi: 10.1061/(ASCE)GT.1943-5606.0000462
![]() |
[31] | Wang H, Shen X, Wang X, et al. (2012) Experimental study on mechanical properties and microstructure of cement mortar composite soil. Chin J Rock Mech Eng 31: 3264–3269 (in Chinese). |
[32] |
Farouk A, Shahien M (2013) Ground improvement using soil-cement columns: Experimental investigation. Alex Eng J 52: 733–740. http://dx.doi.org/10.1016/j.aej.2013.08.009 doi: 10.1016/j.aej.2013.08.009
![]() |
[33] |
Gyeongo K, Takashi T, Athapaththu A (2013) Strength mobilization of cement-treated dredged clay during the early stages of curing. Soils Found 2: 375–392. http://dx.doi.org/10.1016/j.sandf.2015.02.012 doi: 10.1016/j.sandf.2015.02.012
![]() |
[34] |
Chian S, Chim Y, Wong J (2016) Influence of sand impurities in cement-treated clays. Géotechnique 67: 31–41. https://doi.org/10.1680/jgeot.15.P.179 doi: 10.1680/jgeot.15.P.179
![]() |
[35] |
Du J, Liu B, Wang Z, et al. (2021) Dynamic behavior of cement-stabilized organic-matter disseminated sand under cyclic triaxial condition. Soil Dyn Earthq Eng 147: 106777. https://doi.org/10.1016/j.soildyn.2021.106777 doi: 10.1016/j.soildyn.2021.106777
![]() |
[36] |
Wahab N, Talib M, Malik A, et al. (2023) Effect of cement stabilized peat on strength, microstructure, and chemical analysis. Phys Chem Earth Parts A/B/C 129: 103348. https://doi.org/10.1016/j.pce.2022.103348 doi: 10.1016/j.pce.2022.103348
![]() |
[37] |
Fatahi B, Le T, Fatahi B, et al. (2013) Shrinkage properties of soft clay treated with cement and geofibers. Geotech Geol Eng 31: 1421–1435. https://doi.org/10.1007s/10706-013-9666-y doi: 10.1007s/10706-013-9666-y
![]() |
[38] |
Du Y, Jiang N, Liu S, et al. (2014) Engineering properties and microstructural characteristics of cement-stabilized zinc-contaminated kaolin. Can Geotech J 51: 289–302. https://doi.org/10.1139/cgj-2013-0177 doi: 10.1139/cgj-2013-0177
![]() |
[39] |
Zhao C, Zhao H, Chang Y, et al. (2014) Experimental study on physicochemical properties of cement-modified strongly expansive soils. J Dalian Univ Technol 6: 604–611 (in Chinese). https://doi.org/10.7511/dllgxb201406002 doi: 10.7511/dllgxb201406002
![]() |
[40] |
Liang S, Zeng W (2018) Experimental study of nansha silt soil reinforced with cement and fly-ash during wetting-drying cycles. Ind Constr 48: 83–86 (in Chinese). https://doi.org/10.13204/j.gyjz201807015 doi: 10.13204/j.gyjz201807015
![]() |
[41] |
Zidan A (2020) Strength and consolidation characteristics for cement stabilized cohesive soil considering consistency index. Geotech Geol Eng 38: 5341–5353. https://doi.org/10.1007/s10706-020-01367-6 doi: 10.1007/s10706-020-01367-6
![]() |
[42] |
Phutthananon C, Jongpradist P, Nakin S, et al. (2022) State parameter governing the mechanical properties of cement-treated clays. Mar Georesources Geotechnol 41: 388–399. https://doi.org/10.1080/1064119X.2022.2049935 doi: 10.1080/1064119X.2022.2049935
![]() |
[43] |
Sai Chandu A, Rama Subba Rao GV (2021) Strength and durability characteristics of red soil stabilised with foundry sand and cement. Arab J Sci Eng 46: 5171–5178. https://doi.org/10.1007/s13369-021-05423-y doi: 10.1007/s13369-021-05423-y
![]() |
[44] |
Xu M, Liu L, Deng Y, et al. (2021) Influence of sand incorporation on unconfined compression strength of cement-based stabilized soft clay. Soils Found 61: 1132–1141. https://doi.org/10.1016/j.sandf.2021.06.008 doi: 10.1016/j.sandf.2021.06.008
![]() |
[45] |
Liang S, Wang Y, Feng D (2023) Experimental study on strength and dry-wet cycle characteristics of south China coastal soft soil solidified by cement collaborating sand particles. Appl Sci 13: 8844. https://doi.org/10.3390/app13158844 doi: 10.3390/app13158844
![]() |
[46] |
Nakayenga J, Mutsuko I, Guharay A, et al. (2023) Effect of limestone and granite stone powder on properties of cement-treated clay composites and their socioeconomic and environmental impacts. Constr Build Mater 393: 132064. https://doi.org/10.1016/j.conbuildmat.2023.132064 doi: 10.1016/j.conbuildmat.2023.132064
![]() |
[47] |
Phutthananon C, Tiyawan N, Jongpradist P, et al. (2023) Investigation of strength and microstructural characteristics of blended cement-admixed clay with bottom ash. Sustainability 15: 3795. https://doi.org/10.3390/su15043795 doi: 10.3390/su15043795
![]() |
[48] |
Chen C, Zhang C, Liu X, et al. (2023) Effects of freeze-thaw cycles on permeability behavior and desiccation cracking of dalian red clay in China considering saline intrusion. Sustainability 15: 3858. https://doi.org/10.3390/su15043858 doi: 10.3390/su15043858
![]() |
[49] |
Yarba N, Kalkan E, Akbulut S (2007) Modification of the geotechnical properties, as influenced by freeze-thaw, of granular soils with waste additives. Cold Reg Sci Technol 48: 44–54. https://doi.org/10.1016/j.coldregions.2006.09.009 doi: 10.1016/j.coldregions.2006.09.009
![]() |
[50] |
Liu J, Wang T, Tian Y (2010) Experimental study of the dynamic properties of cement- and lime-modified clay soils subjected to freeze–thaw cycles. Cold Reg Sci Technol 61: 29–33. https://doi.org/10.1016/j.coldregions.2010.01.002 doi: 10.1016/j.coldregions.2010.01.002
![]() |
[51] |
Sagidullina N, Abdialim S, Kim J, et al. (2022) Influence of freeze-thaw cycles on physical and mechanical properties of cement-treated silty sand. Sustainability 14: 7000. https://doi.org/10.3390/su14127000 doi: 10.3390/su14127000
![]() |
[52] |
Ma Z, Xing Z, Zhao Y, et al. (2023) Dynamic strength characteristics of cement-improved silty clay under the effect of freeze-thaw cycles. Sustainability 15: 3333. https://doi.org/10.3390/su15043333 doi: 10.3390/su15043333
![]() |
1. | Nicola De Nitti, Enrique Zuazua, On the Controllability of Entropy Solutions of Scalar Conservation Laws at a Junction via Lyapunov Methods, 2023, 51, 2305-221X, 71, 10.1007/s10013-022-00598-9 | |
2. | Ulrik S. Fjordholm, Markus Musch, Nils H. Risebro, Well-Posedness and Convergence of a Finite Volume Method for Conservation Laws on Networks, 2022, 60, 0036-1429, 606, 10.1137/21M145001X | |
3. | P. Cardaliaguet, N. Forcadel, Microscopic Derivation of a Traffic Flow Model with a Bifurcation, 2024, 248, 0003-9527, 10.1007/s00205-023-01948-8 | |
4. | Pierre Cardaliaguet, Nicolas Forcadel, Régis Monneau, A class of germs arising from homogenization in traffic flow on junctions, 2024, 21, 0219-8916, 189, 10.1142/S0219891624500073 | |
5. | Alexandre M. Bayen, Alexander Keimer, Nils Müller, A proof of Kirchhoff's first law for hyperbolic conservation laws on networks, 2023, 18, 1556-1801, 1799, 10.3934/nhm.2023078 | |
6. | Dilip Sarkar, Shridhar Kumar, Pratibhamoy Das, Higinio Ramos, Higher-order convergence analysis for interior and boundary layers in a semi-linear reaction-diffusion system networked by a k -star graph with non-smooth source terms, 2024, 19, 1556-1801, 1085, 10.3934/nhm.2024048 | |
7. | Taras Mel’nyk, Christian Rohde, Asymptotic expansion for convection-dominated transport in a thin graph-like junction, 2024, 22, 0219-5305, 833, 10.1142/S0219530524500040 | |
8. | Jon Asier Bárcena-Petisco, Márcio Cavalcante, Giuseppe Maria Coclite, Nicola De Nitti, Enrique Zuazua, Control of hyperbolic and parabolic equations on networks and singular limits, 2024, 0, 2156-8472, 0, 10.3934/mcrf.2024015 |
Example 7.3 | Example 7.5 | Example 7.4 | Example 7.1 | Example 7.2 | ||||||
Grid level | EOC | EOC | EOC | EOC | EOC | |||||
3 | 0.10877 | - | 0.11630 | - | 0.14459 | - | 0.07087 | - | 0.09904 | - |
4 | 0.05496 | 0.98 | 0.07136 | 0.70 | 0.08016 | 0.85 | 0.0546 | 0.38 | 0.04913 | 1.01 |
5 | 0.03649 | 0.59 | 0.04372 | 0.71 | 0.04651 | 0.79 | 0.03117 | 0.81 | 0.02844 | 0.79 |
6 | 0.02629 | 0.47 | 0.02255 | 0.96 | 0.02711 | 0.78 | 0.01903 | 0.71 | 0.01627 | 0.81 |
7 | 0.01830 | 0.52 | 0.01360 | 0.73 | 0.01495 | 0.86 | 0.01115 | 0.77 | 0.00919 | 0.82 |
8 | 0.01255 | 0.54 | 0.00653 | 1.06 | 0.00925 | 0.69 | 0.00644 | 0.79 | 0.00527 | 0.80 |
9 | 0.00883 | 0.51 | 0.00325 | 1.01 | 0.00480 | 0.95 | 0.00330 | 0.96 | 0.00268 | 0.98 |
10 | 0.00625 | 0.50 | 0.00160 | 1.02 | 0.00295 | 0.70 | 0.00173 | 0.93 | 0.00150 | 0.84 |
11 | 0.00442 | 0.50 | 0.00086 | 0.90 | 0.00152 | 0.96 | 0.00085 | 1.03 | 0.00084 | 0.84 |
12 | 0.00312 | 0.50 | 0.00040 | 1.10 | 0.00081 | 0.91 | 0.00042 | 1.02 | 0.00047 | 0.84 |
Example 7.3 | Example 7.5 | Example 7.4 | Example 7.1 | Example 7.2 | ||||||
Grid level | EOC | EOC | EOC | EOC | EOC | |||||
3 | 0.10877 | - | 0.11630 | - | 0.14459 | - | 0.07087 | - | 0.09904 | - |
4 | 0.05496 | 0.98 | 0.07136 | 0.70 | 0.08016 | 0.85 | 0.0546 | 0.38 | 0.04913 | 1.01 |
5 | 0.03649 | 0.59 | 0.04372 | 0.71 | 0.04651 | 0.79 | 0.03117 | 0.81 | 0.02844 | 0.79 |
6 | 0.02629 | 0.47 | 0.02255 | 0.96 | 0.02711 | 0.78 | 0.01903 | 0.71 | 0.01627 | 0.81 |
7 | 0.01830 | 0.52 | 0.01360 | 0.73 | 0.01495 | 0.86 | 0.01115 | 0.77 | 0.00919 | 0.82 |
8 | 0.01255 | 0.54 | 0.00653 | 1.06 | 0.00925 | 0.69 | 0.00644 | 0.79 | 0.00527 | 0.80 |
9 | 0.00883 | 0.51 | 0.00325 | 1.01 | 0.00480 | 0.95 | 0.00330 | 0.96 | 0.00268 | 0.98 |
10 | 0.00625 | 0.50 | 0.00160 | 1.02 | 0.00295 | 0.70 | 0.00173 | 0.93 | 0.00150 | 0.84 |
11 | 0.00442 | 0.50 | 0.00086 | 0.90 | 0.00152 | 0.96 | 0.00085 | 1.03 | 0.00084 | 0.84 |
12 | 0.00312 | 0.50 | 0.00040 | 1.10 | 0.00081 | 0.91 | 0.00042 | 1.02 | 0.00047 | 0.84 |