1.
Introduction
Singularly perturbed problems (SPPs) frequently appear in several fields of applied mathematics, such as fluid dynamics [22], chemical reactor theory [26], etc. These problems are characterized by the presence of a small positive parameter multiplying the highest-order derivative, but other small parameters may appear affecting other terms. Depending on the parameter location and the smoothness of the coefficients of the SPP, the solution shows steep gradients on the boundary and interior parts of the domain, which are characterized as boundary and interior layers. The layer occurrence causes several hurdles (which include huge computational costs in the case of uniform meshes) in the numerical analysis, which is the reason to consider an adaptive mesh. The simplest adaptive mesh in this context is due to Shishkin [9], who proposed a piecewise uniform mesh to capture the layer behavior. In contemporary literature, numerical methods for SPPs (involving only a diffusion parameter) with smooth data can be seen in [6,9,20], while for non-smooth data, the works [1,7,8,10] considered Shishkin meshes.
Two-parameter problems involving convection (εc) and diffusion (εd) parameters extend the convection and reaction-dominated models. In recent years, several higher-order accurate boundary layer resolving numerical methods based on hybrid schemes, cubic spline schemes, asymptotic expansion methods, etc., have been presented in [11,12,15,24,25,27] for two-parameter problems with smooth data. The non-smooth data produce interior layers in addition to the boundary layers, whose sharpness depends on the sign of the convection coefficient and has been investigated by various researchers for both singularly perturbed ordinary and partial differential equations (for instance, see [2,3,4,17,19,21]). The works [7,10,14] (involving only a diffusion parameter) are devoted to the numerical analysis of interior layers due to the presence of a discontinuous convection coefficient. A rigorous analysis of the effect of all possible subclasses of discontinuous convection coefficients can be found in Riordan [14,18]. These above works motivated us to develop a higher-order numerical approximation for a two-parameter singularly perturbed problem where some of the coefficients have jump discontinuities.
Motivated by the studies in [14,18], we consider the following two-parameter singularly perturbed problem with a discontinuous convection coefficient and a discontinuous source term:
Here εd and εc are known as singular perturbation parameters, where 0<εd≪1, 0≤εc≤1. For simplicity, we consider the domain as ¯Γ=[0,1], with Γ=(0,1), Γ−=(0,d) and Γ+=(d,1). Here b(x) is assumed to be a sufficiently smooth function in ¯Γ satisfying b(x)≥β>0 and a(x), f(x) are sufficiently smooth in (Γ−∪Γ+)∪{0,1}. Also, a(x) and f(x) have a jump discontinuity at d∈Γ, where the jump of ω(x) at x=d is denoted as [ω](d)=ω(d+)−ω(d−). These assumptions ensure that the SPP (1.1)-(1.2) has a solution u(x)∈C0(¯Γ)∩C1(Γ)∩C2(Γ−∪Γ+).
Note that εc=1 reduces the general problem in (1.1) to a convection-diffusion problem [7], and εc=0 reduces (1.1) to a reaction-diffusion problem [8]. The nature of the solution behaves differently with respect to the ratio of the parameters εd/ε2c. The solution behaves similarly to a dissipative case, when εd/ε2c→0 as εc→0, and it acts as in the dispersive case when ε2c/εd→0 as εd→0 [16]. Hence, we consider the following two cases for the numerical analysis of (1.1):
Case (i): √αεc≤√γεd,
Case (ii): √αεc≥√γεd,
where γ=min¯Γ{b(x)α(x)}, where α(x)=α1,x<d, and α(x)=α2,x>d.
Apart from the assumptions considered in (1.3), we have also noted the case when the sign of the convection coefficient is reversed in Γ− and Γ+ (i.e., a(x)≥α1>0,x∈Γ− and a(x)≤−α2<0,x∈Γ+) as a remark.
For two-parameter singularly perturbed problems with smooth data, Riordan et al. [15] analyzed the upwind scheme on a Shishkin mesh and obtained a uniform accuracy of O(N−1ln2N), where N defines the number of partitions in the domain. Later, Gracia et al. [11] established a higher-order numerical method for this problem, which is based on the combination of upwind, central difference, and mid-point schemes in various partitions of the domain. Their numerical method provided parameter-uniform convergence of accuracy O(N−2ln3N), if √αεc≤√γεd, and O(N−2ln2N), if √αεc≥√γεd on a Shishkin mesh. In [23], the authors introduced a discontinuity in the source term for this problem and obtained O(N−1ln2N), for √αεc≤√γεd and O(N−1ln3N), for √αεc≥√γεd on a Shishkin mesh. Several other adaptive meshes, based on arc-length equidistribution [13] and curvature-based monitor functions [5], have also been considered to improve uniform accuracy up to the first-order.
The above articles motivated us to develop a higher-order numerical analysis for two-parameter problems with a discontinuous convection coefficient and source term. The discontinuous data make the numerical analysis different as it gives rise to the interior layer in addition to the boundary layer. We consider the error analysis on the Shishkin mesh and show that the error is independent of the convection and diffusion parameters.
Throughout this article, we denote C as a generic positive constant independent of the number of nodal points and the perturbation parameters εd, εc. The convergence is estimated in the infinity norm, which is denoted as ‖u‖Ω=maxx∈Ω|u(x)| for a function u(x) defined on a general domain Ω. We also write ‖.‖=‖.‖Ω, if the norm and domain are obvious. Accordingly, the corresponding discrete norm is denoted as ‖.‖=‖.‖ΩN. Without loss of generality, we assume that the number of mesh intervals N is divisible by 2.
The paper is arranged as follows: In Section 2, we note the existence of the solution and derive a minimum principle for (1.1)-(1.2), from which it follows the stability of the solution u(x). Some estimates of the solution and its derivatives are also stated here. Section 3 presents a discrete problem based on a hybrid finite difference scheme corresponding to the continuous problem. A decomposition of the discrete solution is introduced in Section 4, which helps us evaluate the truncation error estimate. In Section 5, we have shown that this estimate provides a higher-order εd-εc uniform numerical approximation in the discrete maximum norm. Numerical examples, given in Section 6, validate the theoretical findings. In the end, we draw a conclusion by highlighting the major contribution of the paper.
2.
Derivative bounds of the continuous solution
In this section, we consider a few analytical properties of the solution of (1.1)-(1.2). The solution is decomposed into regular and singular components, which describe the solution behavior at the outer and inner regions of the boundary and interior layers, respectively. This decomposition will be used in the subsequent sections to obtain a parameter-uniform error estimate. We begin this section with an existence theorem.
Theorem 2.1. The SPP (1.1)-(1.2) has a solution u(x), which belongs to the class C0(¯Γ)∩C1(Γ)∩C2(Γ−∪Γ+).
Proof. By using the constructive method presented in [8,10], a solution of the SPP (1.1)-(1.2) can be obtained. □
The operator L at (1.1) satisfies the following minimum principle on ¯Γ.
Lemma 2.1. If a function u(x)∈C0(¯Γ)∩C2(Γ−∪Γ+) satisfies u(0)≥0, u(1)≥0, Lu(x)≤0 for all x∈(Γ−∪Γ+) and [u′](d)≤0, then u(x)≥0∀x∈¯Γ.
Proof. For the proof and the existence of the solution of the SPP (1.1)-(1.2), the reader is referred to [18]. □
As a consequence, we get the following stability estimate:
Lemma 2.2. Let u(x) be a solution of (1.1)-(1.2), then
One can also obtain the following bounds for the solution derivatives when a(x) and f(x) have a jump discontinuity at x=d (see [18]).
Lemma 2.3. Let u(x) be the solution of problem (1.1)-(1.2), where |u(0)|≤C and |u(1)|≤C. Then, for all 0≤k≤4, it holds
Proof. Let us begin the discussion on the domain Ω−. Consider any point x∈(0,d) and a neighborhood Np=(p,p+r), where r is positive, such that x∈Np⊂(0,d). Since u is differentiable in Np, the Mean Value Theorem implies that there exists q∈Np such that u′(q)=u(p+r)−u(p)r. Now, we obtain
We can define u′(x) as
Further, the result for the first derivative can be deduced assuming that x−q≤r and r=√εd, from which we obtain
The bounds for the second derivative will be
Now, we differentiate (1.1) to obtain the third derivative bounds, which result in
Finally, we derive the bounds for the fourth derivative as follows:
After simplifying it, we can write the required result as
The proof for the domain Ω+ can be obtained using similar arguments. □
Before going into further details about the decomposition of u(x) into regular and singular components, we will consider the following: Let F(x) be a smooth function in (Γ−∪Γ+), such that F(x) and its derivatives have a jump discontinuity at d∈Γ. Consider u(x)∈C1(Γ)∩C2(Γ−∪Γ+), such that
It can be proven that problem (2.1) has a unique solution [7]. Let
and further let u∗l(x) be the solution of
Similarly, one can define u∗r(x) on the interval [d,1]. Now
To establish a sharper bound on the error analysis, the solution u(x) is decomposed into a regular component v(x) and a singular component w(x) such that
and w(x)=wl(x)+wr(x) where
We now define the regular and singular components as the solutions to the following problems, respectively.
and
Now, let us consider the case (i): √αεc≤√γεd.
We first define v0(x),v1(x) and v2(x) as the solutions to the following problems:
Choose v3(x)∈C0(¯Γ)∩C1(Γ)∩C2(Γ−∪Γ+), such that
Adopting the procedure from [11,18], we obtain the upper bounds of the derivatives of the regular and the singular components given in the following Lemmas 2.4 and 2.5.
Lemma 2.4. When √αεc≤√γεd, the regular component v(x) and its derivatives satisfy the following bounds:
Lemma 2.5. When √αεc≤√γεd, the singular components wl(x) and wr(x) and their derivatives satisfy the bounds
where
Now consider the case (ii): √αεc≥√γεd.
Let v0(x),v1(x), and v2(x) be the solutions of the following problems, respectively:
Choose v3(x)∈C0(¯Γ)∩C1(Γ)∩C2(Γ−∪Γ+) such that
Using the reasoning given in [11,18], the following Lemmas 2.6 and 2.7 can be proved for the case √αεc≥√γεd.
Lemma 2.6. When √αεc≥√γεd, the regular component v(x) satisfies the following bounds
Lemma 2.7. When √αεc≥√γεd the singular components wl(x) and wr(x) satisfy the bounds
where
It can be verified that v(x)+wl(x)+wr(x) satisfies the problem (1.1)-(1.2). Therefore, the unique solution to the problem is
3.
Discretization of the problem using a Hybrid scheme
In this section, we introduce a difference scheme to discretize the continuous problem (1.1)-(1.2). The discrete problem combines the standard upwind, mid-point, central difference, and five-point difference schemes. The five-point difference scheme is applied at the point of discontinuity to ensure higher-order (in this case, second-order) accuracy. This scheme is reduced to a three-point structure to preserve the monotonicity property with almost second-order accuracy. The numerical scheme, obtained by combining all these operators, maintains the monotonicity property.
The discrete problem will be defined on an a priori adaptive piecewise uniform mesh, which is dense inside the boundary and interior layer regions. To construct this mesh, we first divide the domain ¯Γ into six subintervals:
for some τ1,τ2,τ3, and τ4. The mesh points are denoted as ¯ΓN={xi}N0, where xN/2 denotes the point of discontinuity, xN/2=d. On these mesh points, we define the discrete solution as Ui. The transition parameters τ1,τ2,τ3, and τ4 in ¯Γ are chosen as follows:
where θ1 and θ2 are defined in the previous section. Now we construct a uniform mesh on each of the subintervals [0,τ1],[d−τ2,d],[d,d+τ3] and [1−τ4,1], so that each one contains N/8+1 uniform mesh points, and the subintervals [τ1,d−τ2] and [d+τ3,1−τ4] contain N/4+1 uniform mesh points respectively. The mesh sizes in each of the subintervals from left to right of ¯Γ are denoted as h1=8τ1/N,h2=4(d−τ2−τ1)/N,h3=8τ2/N,h4=8τ3/N,h5=4(1−d−τ3−τ4)/N,andh6=8τ4/N. On the above adaptive mesh ¯ΓN, we discretize the BVP (1.1)-(1.2) as
The discretization (3.2), based on the upwind scheme, is almost first-order accurate [23] (see the numerical Section 6). Hence, our goal is to improve the order of accuracy. Now, we construct an almost second-order accurate hybrid scheme for (1.1)-(1.2) by combining the following central, upwind, and mid-point difference schemes with a five-point scheme:
where, D0Ui=Ui+1−Ui−1hi+hi+1,D+Ui=Ui+1−Uihi+1, D−Ui=Ui−Ui−1hi, δ2Ui=1¯hi(D+Ui−D−Ui), ¯zi=zi+zi+12 and D∗={D−,ifi<N/2,D+,ifi>N/2.
At xN/2=d, we use a five-point difference scheme by combining the second-order accurate one-sided forward difference approximation u′(x)≈(−3U(x)+4U(x+h3)−U(x+2h3))/2h3 with the backward difference approximation u′(x)≈(3U(x)−4U(x−h4)+U(x−2h4))/2h4. At the point of discontinuity, we define the scheme as:
where h=max{h3,h4}.
The finite difference scheme, which involves the central, upwind, mid-point, and five-band difference schemes on the piecewise uniform mesh, will be written as
where LN is defined as follows:
When √αεc≤√γεd,
and for √αεc≥√γεd
At the transition points τ1 and (d+τ3), the scheme is defined as
At the transition points (d−τ2) and (1−τ4), we define the scheme as
Now, we define the discrete scheme as
The matrix associated with (3.4) does not satisfy the M-matrix condition at the point of discontinuity xi=d. But, without loss of generality, we can convert this five-point difference scheme into a three-point difference scheme (say, LNTUi) by estimating UN/2−2,UN/2+2 from LNcUi so that the new equations have the monotonicity property. To do this, we take
Now we replace the above expressions of UN/2−2,UN/2+2 at the five-point difference scheme (LNtUN/2) to construct a three-point scheme (LNTUN/2) that preserves the monotonicity property and leads to a higher-order accuracy at the point of discontinuity, as given by
So, the reformulated discrete operator (say LN∗Ui) of (3.4) can be written as
where
and
The entries of the stiffness matrix corresponding to LN∗ are given by
Now, we state a few conditions that are used to preserve the monotonic properties of the discrete problem (3.5). In (0,τ1) and (d,d+τ3), note that
For √αεc≤√γεd in (d−τ2,d) and (1−τ4,1), we get
For √αεc≥√γεd in (d−τ2,d) and (1−τ4,1), we get
At xi=d, we have
We note that in order to guarantee that the operator LN∗ is monotone, it is necessary to impose the following assumption:
which will be evident in the proof of the following lemma. Thus, the discrete problem is
The following lemma shows that the discrete operator LN∗ has stability properties analogous to those of the continuous operator L.
Lemma 3.1. If a mesh function Ui satisfies U0≥0, UN≥0, LN∗Ui≤0 for all i=1,2,…,N−1 and LNTUN/2≤0, then Ui≥0 for all i=0,1,…,N.
Proof. The system LN∗Ui=QN∗fi for all 1≤i≤N−1 is a linear system of N−1 equations. We show that the corresponding matrix is diagonally dominant. To show this, it is sufficient to check that the conditions
are fulfilled for the operators defined in (3.11).
For √αεc≤√γεd, the central difference operator (LNc) is used in all the subintervals of ¯ΓN. Therefore, s−i>0 is guaranteed in the subintervals (d−τ2,d) and (d,d+τ3) by the definition of h3 and h4. In the remaining regions, the LNc operator is used when εch1‖a‖,εch2‖a‖,εch5‖a‖,εch6‖a‖<2εd, and thus, s−i>0 is satisfied from (3.7), (3.8), and (3.10).
For √αεc≥√γεd, the central difference operator is applied in (d−τ2,d) and (d,d+τ3). Here, h3 and h4 are given by s−i>0 and s+i>0. The mid-point operator is applied to the layer region (0,τ1). Here, the inequalityhis−i=εd¯hi−εcai>0 is guaranteed since εch1‖a‖<εd2 and s+i=εdhi+1¯hi−bi+12>0 is satisfied if ‖b‖h21<εd4. The inequalities (3.6) and (3.10) make it obvious. The sign pattern of s+i+sci+s−i<0 is direct using the condition sci=−s+i−s−i−¯bi. When the mid-point operator is used for the interval (1−τ4,1) the condition (3.12) is satisfied if s+i=εc¯aihi+1−bi2>0. This is guaranteed since ‖b‖h6<εcα/2.
In the coarse mesh region (τ1,d−τ2) and (d+τ3,1−τ4), s+i>0 also follows from the mid-point operator since 2‖b‖h2<εcα and 2‖b‖h5<εcα on each of the intervals. If 2‖b‖h2,2‖b‖h5≥εcα, the upwind operator LNu is used to preserve s+i>0.
At the point of discontinuity, xN/2=d, where √αεc≤√γεd, s+N/2=2εd+2h4εcaN/2−bN/2h24>0 is guaranteed since ‖b‖h4≤2εcα. For s−N/2=2εd−bh23−2εch3aN/2>0 is assured with ‖b‖h23<εd/4 and εch3‖a‖<εd/2. Again, scN/2=−s−N/2−s+N/2−2bN/2−1[h232εd+εcaN/2−1h3+h242εd+εcaN/2+1h4] follows, scN/2<0. For √αεc≥√γεd, we have s+N/2>0 and s−N/2>0 since ‖b‖h24<εd/4 and 2‖b‖h3<εcα. Similarly, we can show scN/2<0.
Hence, combining all the above relations defined at various mesh points, it can be noted that the matrix corresponding to LN∗ is an M-matrix, and hence it satisfies the discrete minimum principle. □
Lemma 3.2. If Ui is the solution to (3.11), then
where i=0,1,2,…,N and θ=min{α1/d,α2/(1−d)}.
Proof. Let us define the mesh functions
where M=max{|U0|,|UN|}+1θ‖QN∗fi‖.
It is obvious, Θ±0≥0, Θ±N≥0 and also
Similarly, LNmΘ±i≤0 and LNuΘ±i≤0 for the above values of i.
At the point of discontinuity, it is
Applying Lemma 3.1, it follows that
which leads to the desired bound on Ui. □
4.
Truncation error analysis
We address the error analysis in this section. The nodal error is denoted by ei=Ui−u(xi). To bound the nodal error |ei|, we first decompose the discrete solution (in a similar manner, as was done with the continuous solution) of (3.11) as Ui=Vi+Wl,i+Wr,i. The discrete regular component (Vi) and singular components (Wl,i and Wr,i) are again decomposed to obtain sharper bounds. Let us define the mesh functions V−i and V+i, which approximate Vi on either side of the point xi=d. In addition, we construct the mesh functions W−l,i,W+l,i and W−r,i,W+r,i to approximate Wl,i and Wr,i on the left and right sides of xi=d. Here, W−l,i and W−r,i correspond to the left boundary layer and right interior layer, respectively. Similarly, W+l,i and W+r,i are the solutions of the left interior layer and right boundary layer. These mesh functions are useful to show the convergence of the nodal error |ei| inside and outside the layers.
Let the regular components V−i, V+i are the solutions to the following discrete problems:
Similarly, the discrete problem corresponding to the left singular components W−l,i and W+l,i is defined as follows:
The corresponding right singular components W−r,i and W+r,i can be described in a similar way.
Consequently, the solution of the discrete problem (3.11) can be written as
The following lemma provides the bounds for the singular components.
Lemma 4.1. The bounds of the left and right singular components W−l,i,W+l,i,W−r,i and W+r,i are as follows:
where k1=i, k2=N/2+i, and C is a positive constant independent of εc,εd.
Proof. Define the barrier functions
where
and also
Now we will prove that LN∗ψ−l,i≤0,LN∗ψ−r,i≤0. Applying the discrete operator (3.11) on ψ−l,i, we have
Again, for the central difference operator, it follows that
Applying (4.1) in the above equation for both the cases √αεc≤√γεd and √αεc≥√γεd, we obtain
In a similar way, the upwind and mid-point difference schemes on ψ−l,i lead to
and
Combining all the above inequalities, we can conclude that LN∗ψ−l,i≤0. Now we find that, ϕ−l,0>0,ϕ−l,N/2>0 and LN∗ϕ−l,i<0. Hence, by using Lemma 10 in [11], we can prove that ϕ−l,i≥0. Therefore, |W−l,i|≤Ck1∏j=1(1+θ2hj)−1.
Now consider the right layer barrier function ψ−r,i. Applying the discrete operator defined in (3.11) over ψ−r,i, we obtain
Applying the central difference scheme in the place of LN∗, we get
Using (4.1) for both the cases √αεc≤√γεd and √αεc≥√γεd, we obtain
Similar arguments can be used to show that the upwind difference and mid-point difference satisfy
and
Hence, it is clear that, ϕ−r,0>0, ϕ−r,N/2>0 and LN∗ϕ−r,i<0. Therefore, applying Lemma 10 in [11], we obtain ϕ−r,i≥0 and hence |W−r,i|≤CN/2∏j=k1+1(1+θ1hj)−1. Similarly, we can prove the bounds of W+l,i and W+r,i as i varies from N/2+1 to N−1. □
Now, we examine the truncation errors based on the three different operators. At the transition points, the mid-point scheme provides better accuracy in the convective terms compared to the central difference scheme on a non-uniform mesh. On a non-uniform mesh, we have
where ¯hi=(hi+hi+1)/2, and on a uniform mesh with step size h, we have
In the following lemma, we present the local error of the regular component.
Lemma 4.2. The regular component of the truncation error satisfies the following estimate:
Proof. For both the cases √αεc≤√γεdand√αεc≥√γεd, when the mesh is uniform, i.e., τ1=τ2=τ3=τ4=1/8, we have for 1≤i≤N/2,
Similarly, we obtain
When the mesh is non-uniform, we have
At the transition points, the upwind difference scheme is used if 2‖b‖hi≥εcα for i=2,3,5,6; otherwise, the mid-point difference scheme will be used. For both of these schemes, we obtain
Define the barrier function ψi=CN−2ζ(xi)±(Vi−v(xi)), where
Then, ψi satisfies
and D0ψi≤0, D+ψi≤0. Now, applying Lemma 3.1, we obtain
Therefore, we get the required result. □
The subsequent lemma provides the local error of the left and right singular components Wl,i and Wr,i associated with the boundary and interior layers.
Lemma 4.3. Let us assume (3.10). Then, the left and right singular components of the error satisfy the following estimates:
Proof. Firstly, consider the uniform mesh, i.e., τ1=τ2=τ3=τ4=1/8, and for √αεc≤√γεd for i=1,⋯,N/2−1, we have
Similarly, |LN∗(W+l,i−w+l(xi))|≤C(N−1lnN)2∀N/2+1≤i≤N−1.
Again when √αεc≥√γεd, the above inequalities lead to
Similarly, |LN∗(W+l,i−w+l)(xi)|≤CεcN−2ln3N.
Now, we consider the error analysis on the non-uniform mesh. In the left boundary layer region (0,τ1), the truncation error is
If √αεc≤√γεd, from (4.2) we obtain
Similarly, we can obtain the result |LN∗(W+l,i−w+l(xi))|≤C(N−1lnN)2 on the left interior layer region (d,d+τ3).
If √αεc≥√γεd, we obtain from (4.2),
Now consider the barrier function on [0,τ1], see [11]
It is easy to check that the barrier function Ψi is non-negative at the boundary points of [0,τ1] and LNcΨi<0. Hence, from Lemma 10 of [11], it follows Ψi≥0 for all [0,τ1]. Therefore
Similarly, we obtain |W+l,i−w+l(xi)|≤CN−2ln3N for xi∈(d,d+τ3).
When xi∈[τ1,d) for √αεd≤√γεd and √αεd≥√γεd we obtain the truncation error
Similarly, we have |LN∗(W+l,i−w+l(xi))|≤CN−2 for xi∈[d+τ3,1). Following the above procedures, we can prove the bound for the right singular components. □
The above estimates provide the truncation errors at the boundary and interior layer regions (Γ−∪Γ+) except at the point of discontinuity. Estimating the error at the point of discontinuity is not straightforward. The following lemma gives an estimate for the local error at the point of discontinuity.
Lemma 4.4. At the point of discontinuity xN/2=d, the error ed satisfies the following estimate:
Proof. Consider the case √αεc≤√γεd. Then
For √αεc≥√γεd, we have
where we have used Lemma 2.3 and a similar procedure adopted in [1]. □
Remark. When the sign of the discontinuous convection coefficient a(x) is reversed, the above result at the point of discontinuity xi=d takes the form
5.
Error estimate
This section presents the main contribution of the paper. The following theorem provides the εd–εc uniform higher-order error estimate of the computed solution.
Theorem 5.1. Let u(xi) be the solution of the continuous problem (1.1)-(1.2) and Ui be the solution of the discrete problem (3.11) at x=xi. Then, for sufficiently large N satisfying the stability condition (3.10), we have
Proof. From the results of Lemmas 2.3, 4.2, 4.3 and using the procedure adopted in [11], it follows that
The presence of the discontinuity leads to the interior layers in a neighborhood of point d. We consider the following two cases to prove the error at the discontinuity point:
Case (i): √αεc≤√γεd, define the discrete barrier function Φi to be the solution of the problem
where
A straight-forward application of the discrete minimum principle on the intervals [0,d] and [d,1] leads to 0≤Φi≤1. Also
Now, define the ancillary continuous functions u1(x) and u2(x) by
It is to be observed that the solutions of the above equations are
where
Define
Now
Hence, following the analysis given in [8] on the interval [0,d] and [d,1], we obtain
and at the point of discontinuity, i.e., for i=N/2,
Now, define the mesh function
It can be clearly checked that y1(0),y1(1)≥0,LN∗y1(xi)≤0∀xi∈¯ΓN and LNty1(xN/2)≤0. Hence, Lemma 3.1 implies that y1(xi)≥0∀xi∈¯ΓN. Therefore
Case (ii): √αεc≥√γεd. To show the error on the discontinuity point, we define the mesh function y2(xi)=˜Wi±ei on the mesh points in (d−τ2,d+τ3) where,
Applying the central difference operator inside the domains (d−τ2,d] and [d,d+τ3), we obtain
Also,
Hence, we have
Now, y2(d−τ2)≥0,y2(d+τ3)≥0, and LN∗y2(xi)≤0forxi∈(d−τ2,d+τ3). Applying Lemma 3.1 to y2(xi) in the above domain, we get
Hence, combining (5.1)–(5.3), we obtain the desired result. □
6.
Numerical examples
This section experimentally demonstrates the applicability of the hybrid scheme (3.11) and compares it with the existing upwind scheme (3.2) for two test problems. These problems have a jump discontinuity in the convective coefficient and source term. For these problems, the signs of the convection coefficients are different to show different boundary and interior layers. The numerical experiments are conducted on piecewise uniform Shishkin mesh, which changes with the various convection coefficients.
Example 6.1. Consider the two-parameter problem (1.1)-(1.2) with the following discontinuous convection coefficient and source term:
Example 6.2 Consider the singularly perturbed two-parameter BVP (1.1)-(1.2) with the following discontinuous convection coefficient and source term:
As the exact solutions of Examples 6.1 and 6.2 are unknown, we use the double mesh principle [5] to calculate the maximum pointwise error and the corresponding order of convergence of the numerical solution provided by the scheme (3.11). The error ENεd,εc and corresponding order of convergence ρNεd,εc are computed as follows:
Here UNi denotes the numerical solution obtained with N number of mesh intervals, and U2Ni denotes the solution on 2N number of mesh intervals obtained by bisecting the previous original mesh. Similarly, we find the error EN and order of convergence ρN of the existing upwind scheme (3.2) for a fixed value of εc and various values of εd, taken from the set S={εd|εd=10−2,10−4,⋯,10−14}:
Here, the numerical experiments are performed by choosing the constant values α = 1.25, β = 1.0 and γ = 0.8 for Example 6.1 and α = 2.0, β = 1.0, and γ = 0.5 for Example 6.2.
We present the maximum pointwise errors ENεd,εc for Examples 6.1 and 6.2 in Tables 1 and 2, respectively. The corresponding orders of convergence ρNεd,εc for Examples 6.1 and 6.2 are shown in Tables 3 and 4, respectively. These tables show that the order of convergence is almost second-order O(N−2ln2N) when √αεc≤√γεd and O(N−2ln3N) when √αεc≥√γεd for the above two examples. Table 5 presents the maximum error EN and order of convergence ρN using the standard upwind scheme (3.2). This table shows that the existing scheme gives first-order parameter uniform convergence, while Tables 3 and 4 show almost second-order convergence using the hybrid difference scheme. Note that the errors are also lower in Tables 1 or 2 compared to Table 5.
Furthermore, we have plotted the numerical solution and the corresponding error for Example 6.1 in Figures 1 and 2. This shows that the interior layer (due to the discontinuity of the convection coefficient and source term) becomes sharper as εc decreases from 10−2 to 10−4 for fixed εd=10−6 (i.e., the case √αεc≤√γεd and √αεc≥√γεd). Similar behavior can be observed for Example 6.2, whose solution and error graph are depicted in Figures 3 and 4. The Loglog plot of the maximum pointwise error for these problems in Figure 5 also demonstrates the uniform convergence of the numerical solution. From this figure we observe that the numerical (hybrid) method provides parameter-uniform convergence of O(N−2ln3N), if √αεc≤√γεd and O(N−2ln2N), if √αεc≥√γεd on Shishkin mesh.
7.
Conclusions
In this paper, an almost second-order uniformly convergent numerical solution is obtained for a two-parameter singularly perturbed problem where the convection coefficient and source term have a jump discontinuity at an interior point of the domain. The present hybrid difference scheme is a combination of upwind, mid-point, and central difference schemes at the interior points with a five-point scheme at the point of discontinuity so that it preserves the monotonicity property on the Shishkin mesh. Both theoretical and computational results based on this scheme produce almost second-order uniform convergence in the discrete maximum norm. Numerical experiments for the test problems validate the theoretical results. As an extension of this work, in the future, we aim to solve the proposed problem by considering a hybrid difference scheme using a Shishkin-Bakhvalov mesh.
Author contributions
All authors contributed equally to the writing of this article. All authors have accepted responsibility for the entire manuscript content and approved its submission.
Use of AI tools declaration
The authors declare they have not used Artificial Intelligence (AI) tools in creating this article.
Acknowledgments
Vellore Institute of Technology, Vellore supported the work of MC under a SEED grant (Sanction Order No. SG20230081).
The authors would like to thank the editor and reviewers for their constructive comments.
Conflict of interest
Higinio Ramos is the(a) Guest Editor of special issue "Numerical Analysis of Differential Equations with Real-world Applications" for AIMS Mathematics. Higinio Ramos was not involved in the editorial review and the decision to publish this article.
The authors declare that there is no conflict of interest regarding the publication of this manuscript.