1. Introduction
The spontaneous emergence of spatial patterns in reaction–diffusion (R–D) equations was first rigorously explained by Turing in his seminal work [34], where he demonstrated that diffusion-driven instability can destabilise homogeneous equilibria and give rise to non-trivial patterns when chemical species diffuse at disparate rates. Since then, R–D equations have become foundational in modelling diverse pattern formation in, for example, biology, chemistry and physics. The most interesting problem is the two-component R–D system with slow-fast diffusion described by

where
$\varepsilon \gt 0$
is sufficiently small;
$x\in \mathbb{R}$
and
$t\geq 0$
denote space and time, respectively;
$\tau$
is the reaction time constant; and
$H_{1}$
and
$H_{2}$
are sufficiently smooth functions. Gray–Scott equation for autocatalytic reactions [Reference Chen and Ward6, Reference Sewalt and Doelman31, Reference Sun, Ward and Russell33, Reference Wei38] and Gierer–Meinhardt (GM) equation for morphogenesis [Reference Chen and Shen8, Reference Iron, Ward and Wei19, Reference Morimoto25, Reference Wei, Winter and Yang40] are the typical examples. The existence and stability of localised patterns in these models, as well as other R–D systems [Reference Byrnes, Carter, Doelman and Liu5, Reference de Rijk, Doelman and Rademacher10, Reference Kong, Ward and Wei22, Reference Qiao and Zhang29, Reference Tzou, Xie, Kolokolnikov and Ward35], have been studied extensively. However, most previous studies focus on spatially homogeneous media. Though this assumption simplifies the underlying mathematical structure, it significantly limits the applicability of the models to realistic heterogeneous environments.
In contrast, heterogeneity is a ubiquitous phenomenon in nature, manifesting in diverse processes such as fluid convection in combustion, heterogeneous porous structures in solute transport, deposition processes and noise effects in biological systems. In recent decades, the impacts of spatial heterogeneity on the existence, stability and bifurcations of localised patterns in R–D equations have attracted significant attention; see, for example, [Reference Avitabile, Brena and Ward2, Reference Bastiaansen, Chirilus-Bruckner and Doelman3, Reference Chen, Li, Shen and Zhang7, Reference Doelman, Van Heijster and Shen15, Reference Knight, Derks, Doelman and Susanto21, Reference Van Heijster, Doelman, Kaper, Nishiura and Ueda36] and the references therein. Notably, heterogeneity breaks the translational invariance of the system, which makes the analysis more challenging.
In this paper, we investigate a singularly perturbed, non-autonomous two-component R–D system on the real line, namely,

where
$U = U(x,t)$
and
$V = V(x,t)$
denote the slow and fast components, respectively. The spatial heterogeneity is introduced through the potential
$W(U, \chi )$
, which is piecewise smooth in
$\chi$
and exhibits a finite jump discontinuity at
$\chi = \pm L$
, where
$\chi =\varepsilon x$
is a slow variable. This structure models an abrupt change in the underlying medium or energy landscape such as a sharp interface between distinct materials or biological domains. More precisely, we consider

with
$W_1(U)\not \equiv W_2(U)$
; that is, the potentials
$W(U, \chi )$
itself has a jump when
$\chi =\pm L$
. The resulting discontinuity in
$W$
breaks the translational invariance and induces spatial pinning of localised structures, a phenomenon that plays a central role in the emergence, stability and persistence of patterns in heterogeneous media.
The functions
$F$
,
$G$
,
$W_1$
and
$W_2$
are assumed to be sufficiently smooth. To proceed with our analysis, the following assumptions are needed:
-
•
$(\text{A}1)$
$G(U, 0, \varepsilon )=G_U(U, 0, \varepsilon )\equiv 0$ ,
$G_V(U,0, \varepsilon )\gt 0$ and
$G_U(U,V, \varepsilon )\gt 0$ for
$U$ ,
$V\gt 0$ .
-
•
$(\text{A}2)$
$F(U, 0)=F_U(U, 0)=F_V(U,0)\equiv 0$ .
-
•
$(\text{A}3)$
$W_1(U)$ has a local minimum at
$U=0$ , that is,
$\frac {\partial }{\partial U} W_1(0)=0$ and
$\frac {\partial ^2 }{\partial U^2} W_1(0)\gt 0$ .
-
• (A4) For each
$u\gt 0$ , there exists a positive homoclinic solution
$(v_{0}(\xi;\;u), q_{0}(\xi;\;u))$ to the following system:
\begin{equation*} \left \{ {\begin{array}{l} v^\prime =q,\\[5pt] q^{\prime }=G(u,v,0), \end{array}} \right .\end{equation*}
$(v_{0}(\xi;\;u), q_{0}(\xi;\;u))$ converges to
$(0,0)$ as
$|\xi |\to \infty$ and satisfies
$q_{0}(0, u)=0$ . That is, for any
$u\gt 0$ , there exists
$v_m \gt 0$ such that
$\int ^{v_m}_0 G(u, s, 0) \,\textrm {d}s =0$ .
Remark 1. If
$W_1(U)=W_2(U)$
, then
$W(U,\chi )\equiv W_1(U)$
, and system (1.2) reduces to the autonomous one

-
• When
$ W^{\prime }_1(U)=\alpha U$ ,
$F(U,V)=-U^{a_1}V^{b_1}$ and
$G(U,V,\varepsilon )=V-U^{a_2}V^{b_2}$ , then system (1.4) is the generalised GM system, and in particular, it reduces to the classical GM equation when
$a_1=0$ ,
$a_2=-1$ and
$b_1=b_2=2$ .
-
• When
$ W^{\prime }_1(U)=\alpha (U-1)$ ,
$F(U,V)=UV^{2}$ and
$G(U,V,\varepsilon )=\beta V-UV^{2}$ , then system (1.4) reduces to the Klausmeier–Gray–Scott system arising in many fields such as vegetative patterns [Reference Siero, Doelman, Eppinga, Rademacher, Rietkerk and Siteur32] and phase transition [Reference Muratov and Osipov26].
The autonomous systems have translational invariance, while this property fails for the heterogeneous ones.
Doelman and Veerman [Reference Doelman and Veerman16] established the criteria for the existence and stability of pinned pulses in the general autonomous system (1.4). These criteria provide a theoretical framework for analysing pulse solutions in autonomous systems. However, the analysis becomes considerably complex for non-autonomous systems owing to the lack of translation invariance in non-autonomous equation (1.2). Thus, the methods developed in [Reference Doelman, Gardner and Kaper12, Reference Doelman and Veerman16, Reference Veerman and Doelman37] for autonomous systems cannot be applied directly.
To address this issue, we extend the methodology developed in [Reference Derks, Doelman, Knight and Susanto11] to the non-autonomous equation (1.2). The approach developed in [Reference Derks, Doelman, Knight and Susanto11] is suitable for analysing pinned pulses in heterogeneous systems. In our current work, by integrating geometric singular perturbation theory (GSPT) [Reference Fenichel17] and the analytical frameworks established in [Reference Doelman, Gardner and Kaper12] and [Reference Veerman and Doelman37], we investigate the formation of pinned pulses in non-autonomous system (1.2). Our analysis shows that equation (1.2) can admit multi-hump pinned pulse solutions. It is also found that the number of humps depends mainly on the spatial heterogeneity length
$2L$
and the properties of the heterogeneous function
$W(U, \chi )$
. In contrast, for the autonomous case, the number of humps is at most two.
A central motivation for the present study is to investigate the GM system with jump-type spatial heterogeneities in its parameters. By introducing the discontinuities in the reaction and diffusion coefficients, the model captures a range of biologically and chemically relevant scenarios, in which the underlying medium exhibits abrupt spatial transitions. In fact, gene regulatory boundaries, cellular niche interfaces and spatially structured micro-environments are typical heterogeneities. Generally, the classical smooth models fail to reflect the discrete, compartmentalised organisation observed in real situations. In gene regulation, threshold-dependent activation of morphogen-responsive genes can lead to sharp on–off domains in gene expression, which is a behaviour that can be effectively modelled via spatially discontinuous reaction terms [Reference Meinhardt and Gierer24]. In developmental biology, niche transitions and tissue compartmentalisation such as those seen in limb formation or Drosophila wing patterning often give rise to abrupt changes in local kinetics and transport rates [Reference Benson, Maini and Sherratt4]. Likewise, heterogeneities in cellular micro-environments, including localised extracellular matrix structures or biochemical barriers, can create sharp spatial gradients in diffusion or reaction properties [Reference Page, Maini and Monk27]. Wei and Winter [Reference Wei and Winter39] provided a rigorous analysis of the GM system with discontinuous diffusion coefficients, where it is shown that such jumps can induce spike pinning, asymmetry and even multistability.
The first main result of this article is the criteria for the existence of pinned pulse solutions to system (1.2). The existence of such pulses is governed by geometric conditions involving the phase space orbits of certain reduced Hamiltonian systems defined by the heterogeneous potential
$W(U, \chi )$
. Depending on the different signs of the points where the pinned pulse enters the heterogeneous region (
$|\chi |\lt L$
), denoted by
$(u_{in},p_{in})$
, and the points where the pinned pulse enters the fast field (
$|\xi |\lt 1/\sqrt {\varepsilon }$
), denoted by
$(u_0, p_0)$
, four distinct cases arise accordingly. Each case corresponds to a different class of orbit structures across the heterogeneous interface (see Figure 3).
Before we state the existence result, we need the following notations.
Notation 1. We refer to the system

the i-th system for convenience.
Notation 2.
Let
$\mathbf{P}$
denote the set of closed orbits of the 2nd system restricted to the region
$u \gt 0$
. Given an energy level
$h \in \mathbf{P}$
, then the Hamiltonian

defines a periodic orbit. Otherwise, if
$h \notin \mathbf{P}$
, then the corresponding trajectory is non-periodic. Moreover, let
$\mathbf{P}_1$
denote the set of orbits of the 2nd system that possess a turning point (i.e.
$p=0$
) in the region
$u \gt 0$
.
Theorem 1. Consider system (1.2) under the assumptions (A1)–(A4), and define

If

admit non-degenerate solutions
$(u_{in}(h), p_{in}(h))$
and
$(u_0(h), p_0(h))$
, respectively, with
$u_0 \gt u_{in}$
, then for sufficiently small
$\varepsilon \gt 0$
, the following statements hold:
-
(i) If
$ p_{0},\, p_{in} \gt 0$ , and there exists
$ h_0 \notin \mathbf{P}$ such that
(1.7)or\begin{align} \mathcal{H}(u_0(h_0),u_{in}(h_0),h_0) = L, \end{align}
$ h_0 \in \mathbf{P}$ such that
(1.8)where\begin{align} \mathcal{H}(u_0(h_0),u_{in}(h_0),h_0)= L\pmod { T}, \end{align}
\begin{equation*} T = 2\int _{u_{l}(h_0)}^{u_{r}(h_0)} \frac {\textrm { d}u}{\sqrt {2W_2(u)+2h_0}}, \end{equation*}
$ u_{l}(h_0)$ and
$ u_{r}(h_0)$ are the intersection points of the periodic orbit with the
$ u$ -axis, then system (1.2) admits pinned pulse solutions.
-
(ii) If
$ p_{0} \lt 0$ and
$ p_{in} \gt 0$ , and there exists
$ h_0 \in \mathbf{P_1}$ such that
(1.9)where\begin{align} \mathcal{H}(u_0(h_0),u_{in}(h_0),h_0) + 2\mathcal{H}(u_t(h_0),u_{0}(h_0),h_0) = L, \end{align}
$ (u_t(h_0), 0)$ represents the turning point of the orbit of the 2nd system, and
$ u_t(h_0) \gt u_0(h_0)$ , or there exists
$ h_0 \in \mathbf{P}$ such that
(1.10)then system (1.2) admits pinned pulse solutions.\begin{align} \mathcal{ H}(u_0(h_0),u_{in}(h_0)) + 2\mathcal{H}(u_r(h_0),u_{0}(h_0))= L\pmod { T}, \end{align}
-
(iii) If
$ p_{0}, p_{in} \lt 0$ , the 1st system has a homoclinic orbit connecting
$(0,0)$ , and there exists
$ h_0 \in \mathbf{P}$ such that
(1.11)then system (1.2) admits pinned pulse solutions.\begin{align} 2T - \mathcal{H}(u_0(h_0),u_{in}(h_0),h_0)= L\pmod { T}, \end{align}
-
(iv) If
$ p_{0} \gt 0$ and
$ p_{in} \lt 0$ , and the 1st system has a homoclinic orbit connecting
$(0,0)$ , and there exists
$ h_0 \in \mathbf{P}$ such that
(1.12)then system (1.2) admits pinned pulse solutions.\begin{align} \mathcal{H}(u_{in}(h_0),u_{l}(h_0),h_0) +\mathcal{ H}(u_0,u_{l}(h_0),h_0)= L\pmod { T}, \end{align}
To determine the sufficient conditions for the (in)stability of pinned pulses, we analyse the point spectrum of the associated variational equation by using the Evans function. By the singularly perturbed structure of the underlying pulse solutions, the Evans function can be decomposed into the product of a fast transmission function and a slow transmission function. These transmission functions correspond to the lower-order eigenvalue problems that govern the leading order dynamics on the fast and slow scales. More precisely, the slow limiting eigenvalue problem is given by

while the fast limiting eigenvalue problem takes the form

Our analysis shows that the fast transmission function possesses a simple zero in the right half of the complex plane, that is, for values of
$\lambda$
with
$\text{Re}\,\lambda \gt 0$
. However, the Evans function remains nonzero at this point due to the presence of a first-order pole in the slow transmission function. This implies that the ‘
$\mathcal{O}(1)$
-eigenvalues’ generated by the fast transmission function are restricted to the stable left-half complex plane, that is,
$\{\lambda \in \mathbb{C}\,|\, \text{Re}\,\lambda \lt 0\}$
.
In autonomous systems, due to the translational invariance, the eigenvalue
$\lambda =0$
remains a root of the fast transmission function for
$0 \lt \varepsilon \ll 1$
. However, for the non-autonomous system (1.2), this conclusion no longer holds. Consequently, the trivial eigenvalue may be perturbed to an ‘
$\mathcal{O}(\varepsilon )$
-eigenvalue’, whose real part is either positive or negative. This subtle perturbation adds a new layer of complexity to the spectral stability problem. To overcome this difficulty, we derive sufficient conditions for instability by analysing the behaviour of the slow transmission function.
Let
$\left (u_{\pm }(\chi , \lambda ), p_{\pm }(\chi , \lambda )\right )$
and
$\left (v_{\pm }(\xi , \lambda ), q_{\pm }(\xi , \lambda )\right )$
denote solutions to the slow and fast limiting eigenvalue problems, respectively, which satisfy

where
$\Lambda _{f}(\lambda )=\sqrt {G_v(0,0,\varepsilon )+\lambda }$
and
$\Lambda _{s}(\lambda )=\sqrt {W^{\prime \prime }_1(0)+\lambda }$
.
Theorem 2.
The pinned pulses of system (1.2) are unstable if either of the following equations admits a root
$\lambda$
with positive real part:

or

where
$u_{+}(0, \lambda )=u_{+}(\chi =0, \lambda )$
,
$p_{+}(0, \lambda )=p_{+}(\chi =0, \lambda )$
, and

where
$E_f(\lambda )$
is defined in (4.25).
The structure of this paper is as follows. In Section 2, we prove Theorem1, which concerns the existence of stationary pinned pulse solutions for the general system (1.2) by GSPT and the technique of matching. Then in Section 3, we apply the results in Section 2 to a non-autonomous GM system to give the explicit existence of this explicit model. Based on the nonlocal eigenvalue problem (NLEP) method, Section 4 contains the analytical framework on the spectral stability for the general system (1.2) by computing the Evans function. These results are stated in Theorem2. Finally, in Section 5, we apply the stability criteria in Section 4 to a non-autonomous GM system to identify the parameter under which the pinned pulses of this model are unstable.
2. Proof of Theorem 1
A stationary pulse solution to (1.2) corresponds to a homoclinic orbit of the following
$4$
-dimensional singularly perturbed ordinary differential system:

where
$u$
and
$v$
represent the slow and fast variables, respectively, the dot indicates the differentiation in
$x$
and
$0\lt \varepsilon \ll 1$
. To analyse the fast dynamics, we introduce a fast scale
$\xi =\frac {x}{\varepsilon }$
to yield

where the prime denotes the differentiation in
$\xi$
. Systems (2.1) and (2.2) are equivalent if
$\varepsilon \neq 0$
.
Systems (2.1) and (2.2) are reversible with respect to

Due to the definition of
$ W(u,\chi )$
given in (1.3),
$ W(u, -\chi ) = W(u, \chi )$
always holds. Together with the fact that
$ F(u,v)$
and
$ G(u,v,\varepsilon )$
do not explicitly depend on
$ x$
, systems (2.1) and (2.2) remain invariant under the reversibility transformation mentioned above. This reversibility plays a crucial role in simplifying the forthcoming analysis on the existence and stability of pinned pulses.
2.1. Fast connections
Taking the singular limit
$\varepsilon \rightarrow 0$
in (2.2) yields the limiting fast system

in which the slow variables
$(u,p)$
serve as parameters.
By the assumption (A1),
$G(u, 0, 0)=0$
holds for certain values of
$u$
. It thus follows that the set

consists of the equilibria of system (2.3). Furthermore,
$\mathcal{M}$
is normally hyperbolic since
$G_v(u,0,\varepsilon )\gt 0$
, which ensures that each equilibrium in
$\mathcal{M}$
is a saddle of system (2.3). Consequently, both the stable and unstable manifolds
$\mathcal{W}^{s}(\mathcal{M})$
and
$\mathcal{W}^{u}(\mathcal{M})$
are
$3$
-dimensional in the
$4$
-dimensional phase space.
System (2.3) is integrable with the Hamiltonian

The assumption (A4) ensures the existence of a homoclinic pulse in system (2.3). Moreover, the stable and unstable manifolds
$\mathcal{W}^{s}(\mathcal{M})$
and
$\mathcal{W}^{u}(\mathcal{M})$
coincide, forming a
$3$
-dimensional homoclinic manifold, which is composed of a two-parameter family of homoclinic orbits (see Figure 1a).

Figure 1. (a) The
$3$
-dimensional homoclinic manifold of the fast limiting system (2.3) at
$\varepsilon =0$
. (b) A transversal intersection between
$\mathcal{W}^{s}(\mathcal{M}_\varepsilon )$
and
$\mathcal{W}^{u}(\mathcal{M}_\varepsilon )$
for
$0\lt \varepsilon \ll 1$
, which is homoclinic to the slow manifold
$\mathcal{M}_\varepsilon$
.
When
$\varepsilon \gt 0$
is sufficiently small, by GSPT,
$\mathcal{M}$
perturbs to a
$2$
-dimensional locally invariant manifold
$\mathcal{M}_\varepsilon$
, which is
$\mathcal{O}(\varepsilon )$
-close to its counterpart. Since
$\mathcal{M}$
is invariant under the flow of (2.2)
$_2$
, it follows
$\mathcal{M}_\varepsilon =\mathcal{M}.$
Similarly, by GSPT again,
$\mathcal{W}^{s/u}(\mathcal{M})$
perturbs to
$3$
-dimensional locally invariant manifolds
$\mathcal{W}^{s/u}(\mathcal{M}_\varepsilon )$
, which are also
$\mathcal{O}(\varepsilon )$
-close to their counterparts. These manifolds consist of the fast stable and unstable fibres with their bases on the slow manifold
$\mathcal{M}_\varepsilon$
. When
$0\lt \varepsilon \ll 1$
,
$\mathcal{W}^{s}(\mathcal{M}_\varepsilon )$
and
$\mathcal{W}^{u}(\mathcal{M}_\varepsilon )$
no longer coincide. Under appropriate conditions, they can intersect transversely, giving rise to a
$2$
-dimensional manifold
$\mathcal{W}^{s}(T_d) \cap \mathcal{W}^{u}(T_o)$
(see Remark2 and Figure 1b).
Here, we want to remark that although the slow variables in system (2.2) are discontinuous due to the term
$ W(u, \chi )$
, with each subregion, the system under consideration becomes autonomous with respect to
$ \chi$
. Thus, within each subregion, the standard framework of GSPT can still be applied to the two autonomous subsystems. At the positions of interfaces
$\chi = \pm L$
, we match the orbits of the two autonomous subsystems in a continuous, differentiable manner; that is, the orbits together with their derivatives are both connected. In this manner, we can still obtain the
$C^1$
-smooth orbits of the full system (2.2) by using GSPT within each smooth region and smooth matching at the position of heterogeneities.
We now employ the adiabatic Melnikov method [Reference Palmer28, Reference Robinson30] to examine the persistence of the 3-dimensional homoclinic manifold, which is homoclinic to
$\mathcal{M}_\varepsilon$
for
$0 \lt \varepsilon \ll 1$
. The analysis reveals that the homoclinic orbit acts as a dynamical bridge that enables the flow originating from the slow unstable manifold of the saddle on
$\mathcal{M}_\varepsilon$
to return to the slow stable manifold, thereby facilitating a dynamical transition.
To facilitate the analysis, we partition the entire spatial interval
$(\!-\infty , +\infty )$
into three subregions, namely, two slow regions
$I_s^- = (\!-\infty , -\frac {1}{\sqrt {\varepsilon }})$
and
$I_s^+ = (\frac {1}{\sqrt {\varepsilon }}, \infty )$
and a fast region
$I_f = [-\frac {1}{\sqrt {\varepsilon }}, \frac {1}{\sqrt {\varepsilon }}]$
. The boundaries of the fast region
$I_f$
are placed in the transition zone, satisfying
$|\xi | \gg 1$
and
$|x| \ll 1$
. In other words, the exact location of
$\partial I_f$
is not critical [Reference Doelman, Gardner and Kaper12, Reference Doelman and Veerman16]. Indeed, the width of
$I_f$
should be chosen appropriately to ensure that
$v_h(\xi )$
is exponentially small everywhere outside
$I_f$
, while
$u_h(\xi )$
remains approximately constant, to the leading order, within the fast region
$I_f$
.
Remark 2. For convenience, we denote by
$\mathcal{W}^{u}(T_o)$
the collection of unstable fibres with base points on the curve
$T_o$
and by
$\mathcal{W}^{s}(T_d)$
the collection of stable fibres with base points on the curve
$T_d$
. These sets intersect forming a
$2$
-dimensional manifold, denoted as
$\mathcal{W}^{s}(T_d)\cap \mathcal{W}^{u}(T_o)$
. In the subsequent analysis, we will prove that

Remark 3. In our setup, the heterogeneity region
$ [-L, L]$
is defined in terms of the super-slow spatial variable
$ \chi = \varepsilon x=\varepsilon ^2 \xi$
, as introduced in equation (1.2). Thus, in terms of the original spatial variable
$ \xi$
, the corresponding heterogeneous region is rescaled to
$ \left [ -\frac {L}{\varepsilon ^2}, \frac {L}{\varepsilon ^2} \right ]$
, which becomes much wider as
$ \varepsilon \to 0$
. On the other hand, the fast region is defined as
$ I_f = \left [ -\frac {1}{\sqrt {\varepsilon }}, \frac {1}{\sqrt {\varepsilon }} \right ]$
. Therefore, we have

In the fast system (2.2), the Hamiltonian
$H(v,q;\;u)$
becomes a slowly varying function. Its derivative with respect to the fast time variable
$\xi$
is given by

The unperturbed homoclinic orbit
$(u^{(0)},p^{(0)},v_{0}(\xi ),q_{0}(\xi ))$
is perturbed to a solution of the full fast system (2.2), denoted by

which is homoclinic to
$\mathcal{M}_{\varepsilon }$
, provided that the Melnikov integral

has a simple zero. In that case, we have
$\gamma _{h}(\xi )\subset \mathcal{W}^{s}(\mathcal{M}_\varepsilon )\cap \mathcal{W}^{u}(\mathcal{M}_\varepsilon )$
. The improper integral is well-defined because
$\gamma _{h}(\xi )$
converges exponentially to
$(0,0,0,0)$
as
$\xi \rightarrow \pm \infty$
.
To leading order,

Since
$v_{ 0}(\xi )$
is even and
$q_{0}(\xi )$
is odd, so
$\big (G(u^{(0)},v_{0}(\xi ),\varepsilon )-G(u^{(0)},v_{0}(\xi ),0)\big )$
and
$\int ^{v_{0}(\xi )}_0 G_u( u^{(0)},s,0)\,\textrm { d}s$
are even functions. Note that
$u^{(0)}$
is assumed to be positive; it then follows from (A1) that
$G_u( u^{(0)},v,0)$
is also positive. Consequently, solving
$\triangle _{I_{f}}H(u^{(0)},p^{(0)})=0$
gives
$p^{(0)}=0$
.
To compute the correction, we expand
$(u_{h}(\xi ), p_{h}(\xi ), v_{h}(\xi ), q_{h}(\xi ))$
in the power of
$\varepsilon$
, namely,

where the initial conditions are
$u_{h}(0)=u^{(0)}$
and
$u_{j}(0)=0$
for all
$j\geq 1$
. Substituting (2.9) into (2.2) gives

Note that the integral term in
$p_{1}(\xi )$
is an odd function since
$v_{ 0}(\xi )$
is an even function; thus, (2.7) becomes

Since
$u_0$
is positive, the integral in (2.10) is also positive by the assumption (A1). Therefore, we conclude that the condition for vanishing Melnikov integral up to
$\mathcal{O}(\varepsilon ^2)$
is

Remark 4. For any orbit

it follows from the reversibility symmetry that

Therefore,
$\xi = 0$
must be an extremum (specifically, a local maximum) of both
$u_{h}(\xi )$
and
$v_{h}(\xi )$
. By Fermat’s Lemma, this yields
$p_{h}(0)=0$
and
$q_{h}(0)=0$
.
Define the Take-off curve
$T_{o}$
and the Touch-down curve
$T_{d}$
on
$\mathcal{M_{\varepsilon }}$
by

with
$p=\frac {1}{\varepsilon }\frac {d\,u}{d\,\xi }$
. By (2.2), we have

where we have used the facts that
$u_{\xi }=\varepsilon p$
and
$p=\mathcal{O}(\varepsilon )$
during
$I_f$
in the first equation, and
$v_0(\xi )$
is an even function with respect to
$\xi$
in the second equation. So, to leading order, the explicit representations of the Take-off curve
$T_{o}$
and the Touch-down curve
$T_{d}$
on
$\mathcal{M_{\varepsilon }}$
are, respectively

where
$\mathcal{T}_{\mp }(u)=\mp \varepsilon \int ^{0}_{-\infty } F(u,v_0(\xi ))\text{ d}\xi$
and
$p=\frac {1}{\varepsilon }\frac {du}{d\xi }$
.
Remark 5. The homoclinic solution
$(v_{0}(\xi;\;u), q_{0}(\xi;\;u))$
of system (2.3) provides the leading order solution of (2.2), given by

which is homoclinic to
$\mathcal{M}_{\varepsilon }$
. Moreover, its asymptotic limits
$\lim _{\xi \to \pm \infty } \phi _h(\xi )$
correspond precisely to the Touch-down and Take-off curves.
2.2. Slow dynamics on the slow manifold
A singular homoclinic orbit is constructed by concatenating trajectories of the fast and slow limiting systems. Though this singular homoclinic orbit is not an exact solution to system (2.2), GSPT ensures that a true homoclinic orbit exists within a small neighbourhood of this singular trajectory. In the previous subsection, the trajectories of the fast limiting system had been derived. In this section, we focus on the slow dynamics on the slow manifold and match them with the fast segments to complete the full orbit.
On the slow manifold

the dynamics of system (2.2) are governed by

In the super-slow coordinate
$\chi =\varepsilon ^2 \xi$
, the slow limiting system can be transformed into

which is a non-autonomous Hamiltonian system with

It follows from (A3) that system (2.18) admits a saddle
$(0,0)$
. The slow trajectories can be constructed via the phase plane analysis on two autonomous Hamiltonian systems,

where
$W^{\prime }_i(u)$
are the potentials outside and inside the heterogeneous region, respectively. Appropriate boundary conditions must be imposed to ensure that the matching between the trajectories and their derivatives at the endpoints of each region can be performed. These trajectories decay exponentially to
$(0,0)$
as
$\chi \rightarrow \pm \infty$
. In this manner, the solutions to system (2.18) are
$\textrm{C}^1$
-smooth.
In the following, we will demonstrate that the value of the Hamiltonian
$h$
inside the heterogeneity with the length
$2L$
serves as an important parameter to characterise the pinned pulses. A pinned pulse is generally determined by four points in the phase plane, namely:
-
• The points where the pinned pulse enters and exits the heterogeneous region, respectively, denoted by
$(u_{in}, p_{in})$ and
$(u_{out}, p_{out})$ .
-
• The points where the pinned pulse enters and exits the fast field, respectively, denoted by
$(u_0, p_0)$ and
$(u_d, p_d)$ .
Since system (2.2) is symmetric under
$\chi \rightarrow -\chi$
, we have

Therefore, only
$(u_{in}, p_{in})$
and
$(u_0, p_0)$
need to be determined. More precisely,
$(u_{in}, p_{in})$
is the intersection between the unstable manifold of the saddle
$(0, 0)$
outside the heterogeneity (the purple solid curve in Figure 2) and the level set
$H_s(u, p) = h$
within the heterogeneous region (the black curve). They satisfy

where

is required to ensure that the intersection between such two slow trajectories is transverse, and
$h$
represents the Hamiltonian value within the heterogeneity. By solving system (2.21), we can obtain
$(u_{in}, p_{in}) = (u_{in}(h), p_{in}(h))$
, where
$h$
acts as a free parameter.

Figure 2. Singular homoclinic orbit of system (2.2) with
$1$
fast and
$4$
slow segments. Left panel: two slow segments outside the heterogeneity are marked by the solid purple curves, indicating the stable and unstable manifolds of the saddle
$(0, 0)$
; two slow segments inside the heterogeneity (the black curves), representing the orbits defined by
$H_s(u,p)=h$
; and a fast segment is indicated by the red dashing line. In the figure, the yellow dot represents
$(u_{in}, p_{in})$
, and the green one stands for
$(u_{out}, p_{out})$
. Right panel: the thick red curve indicates the whole singular homoclinic orbit.
Remark 6. Let
$u=u_{in,k}\gt 0$
,
$k=1, 2, \cdots , K$
, be the
$K$
non-degenerate solutions obtained from (2.21), then there exist
$2K$
distinct points
$(u_{in,k}, p^{\pm }_{in,k})$
. These points play a key role in constructing the transversely intersecting homoclinic orbits.
Based on the curves
$T_{o,d}$
on
$\mathcal{M}_\varepsilon$
defined in (2.14), a homoclinic orbit
$\phi _h$
of the limiting fast system (2.3) can be connected with the slow orbits on the slow manifold
$\mathcal{M}_\varepsilon$
to construct a singular homoclinic orbit. To guarantee that this singular orbit persists under small perturbations, it is essential that the Take-off curve
$T_o$
intersects transversely with the orbit of the 2nd system (2.20). The corresponding intersection point
$(u_0, p_0)$
is governed by

where it is also required that

such that the matching is transverse.
If both equations (2.21) and (2.22) admit non-degenerate solutions
$(u_{in}(h), p_{in}(h))$
and
$(u_0(h), p_0(h))$
, this implies that
-
• The outer and inner slow trajectories intersect transversely at the position of heterogeneity.
-
• The slow trajectory intersects with the fast jump trajectory transversely.
However, the Hamiltonian value
$h$
remains undetermined now. To determine it, we impose the condition that the ‘time’ for the slow trajectory to evolve from
$(u_{in}(h), p_{in}(h))$
to
$(u_0(h), p_0(h))$
is exactly
$L$
(the half-width of the heterogeneous region). Once this condition is satisfied, we get a singular homoclinic orbit from the saddle to itself with transversality. By GSPT, there are true homoclinic orbits in a small neighbourhood of this singular configuration.
Remark 7. The relative positions of
$(u_{in}, p_{in})$
and
$(u_{0}, p_{0})$
, respectively, determined by (2.21) and (2.22) are crucial for constructing the different types of singular homoclinic orbits. It is worth noting that the orbits in the 2nd system possessing through ‘turning points’ are particularly important for constructing pinned pulses with certain dynamical properties.
3. Pinned pulses in a non-autonomous Gierer–Meinhardt equation
Consider

where
$\sigma \neq 0$
,
$d\gt 1$
and

The associated fast limiting system is

which admits a saddle
$(0,0)$
and a homoclinic orbit described by

The explicit homoclinic orbit solution (3.4) of (3.3) is well-known (see, e.g. [Reference Sewalt and Doelman31, Reference Veerman and Doelman37]). By direct calculations, the Take-off and Touch-down curves are, respectively

Note that for
$\sigma \gt 0$
, the Take-off curve lies in the first quadrant of the
$(u, p)$
plane, while for
$\sigma \lt 0$
, it lies in the fourth quadrant. Thus, changing the sign of
$\sigma$
reverses the direction of the fast jump (see Figures 3 and 4).
The limiting slow system corresponding to (2.18) becomes

which has a saddle at
$(0,0)$
when
$\alpha _1\gt 0$
. This is a piecewise Hamiltonian system with

The matching condition for the entry point
$(u_{in}, p_{in})$
is given by

where
$h$
represents the Hamiltonian within the heterogeneous region. Similarly, the matching condition between the slow and fast orbits can be given by

In what follows, we consider two distinct cases.
3.1. Case 1:
$g(\chi )\equiv 0$
.
Through this simplest case, we can illustrate the key idea of the geometric approach clearly.
Under
$g(\chi )\equiv 0$
, the matching conditions in equation (3.8) reduce to

Solving these two equations yields the explicit expressions for
$u_{in}$
and
$p_{in}$
in terms of
$h$
, namely,

where
$\textrm {sign}(h)=\textrm {sign}(\alpha _1-\alpha _2)$
, which ensures that
$u_{in}(h),\, p_{in}(h)\gt 0$
. By direct calculations, when
$\sigma \gt 0$
and
$\textrm {sign}(h)=\textrm {sign}(\alpha _1-\alpha _2)$
, (3.9) can be simplified as

where we only consider
$u_0(h)\in \mathbb{R}_+$
(
$u_{0}(h)=u^-_{0}(h)$
is valid only when
$\alpha _1\lt \alpha _2$
, which is discarded). Thus, the condition
$u_0(h)\gt u_{in}(h)$
is equivalent to

Within the heterogeneous region, the slow orbit satisfies

with
$p=u_{\chi }$
. Therefore, the relationship between the length
$L$
and the Hamiltonian
$h$
is

where

and
$\Delta =\alpha _2h$
. By combining (3.11), (3.12) and (3.14), the value of
$h$
can be implicitly determined. This determines the exact slow orbit used in the construction of the pinned pulse solution.
The unstable and stable manifolds of the saddle
$(0,0)$
are given by

where
$x_*\gt 0$
is introduced such that
$u_{h,-}(\!-L )=u_{in}$
. By the reversible symmetry, it follows that
$u_{h,+}(L )=u_{out}$
. Within the heterogeneous region (
$|x|\lt L$
), the expression of the orbit is given explicitly by

where the constants
$A_{1,2}$
and
$B_{1,2}$
are determined to ensure that
$u_{p,0}(\chi )$
is continuously differentiable at
$\chi =\pm L$
. Thus, the composite solution can be defined by

Hence, the smooth matching condition at
$\chi =\pm L$
is given by

Solving (3.19) directly yields

Theorem 3.
Let
$\varepsilon \gt 0$
be sufficiently small, and
$\alpha _1$
,
$\alpha _2,\,\sigma ,\,L\gt 0$
be given. Assume that (3.13) admits a non-degenerate root
$h$
satisfying
$\textrm {sign}(h)=\textrm { sign}(\alpha _1-\alpha _2)$
and that (3.14) holds, then system (3.1) with
$g(\chi )\equiv 0$
admits a pinned pulse solution
$(u_{h}(\xi ), v_{h}(\xi ))$
with

for
$\xi \in I_f$
, where
$v_{0}(\xi;\;u_0(h))$
and
$u_0(h)$
are given by (3.4) and (3.12), respectively. For
$\xi \in I_s^\pm$
, we have

where
$u_{p,0}(\chi )$
is defined in (3.18).
3.2.
Case 2:
$g(\chi )\neq 0$
.
In the previous subsection, we study the
$g(\chi )\equiv 0$
case; that is, the slow limiting system is piecewise linear. When
$g(\chi ) \neq 0$
, the system becomes piecewise nonlinear, leading to significantly increased complexity.
Outside the heterogeneous region (
$|\chi | \gt L$
), equation (3.6) turns out to be

When
$\alpha _1, \gamma _1 \gt 0$
, this equation admits a homoclinic orbit given by

Therefore, the unstable and stable manifolds of the saddle
$(0, 0)$
outside the heterogeneous region can be, respectively, parameterised by

where
$x_{*} \gt 0$
is chosen such that
$u_{h,\pm }(\chi =\pm L)=u_{in}$
.
We denote the segments of the pinned pulse within the heterogeneous region (
$|\chi |\lt L$
) by
$u_{i,\pm }(\chi ,h)$
, which are the solutions to

If
$u_{i,\pm }(\chi ,h)$
are the homoclinic orbits of the 2nd system, then their expressions are given by

where the constant
$y_{*}\gt 0$
is determined by the matching condition

As a result, equation (3.8) reduces to the algebraic relation

which admits a unique positive solution

provided
$\textrm {sign}(\gamma _1 - \gamma _2) = \textrm {sign}(\alpha _1 - \alpha _2)$
.
Theorem 4.
Let
$\varepsilon \gt 0$
be sufficiently small, and let the parameters
$ \alpha _1, \alpha _2 \gt 0$
,
$ \gamma _1, \gamma _2 \gt 0$
,
$ d \gt 1$
,
$ \sigma \gt 0$
and
$ L \gt 0$
be given. Suppose that equations (3.8) and (3.9) admit non-degenerate solutions
$ (u_{in}(h), p_{in}(h))$
and
$ (u_0(h), p_0(h))$
, respectively, with
$ u_0 \gt u_{in}$
, then the following statements hold:
-
• When
$ p_{in} \gt 0$ .
-
– If
(3.31)admits a non-degenerate root\begin{equation} \int _{u_{in}}^{u_0} \frac {\,\textrm {d}u}{\sqrt {2h + \alpha _2 u^2 - \frac {2\gamma _2}{d+1}u^{d+1}}} = L \end{equation}
$ h \gt 0$ , then system (3.1) admits a pinned pulse solution shown in Figure 3a , b .
-
– If
(3.32)admits a non-degenerate root\begin{equation} \int _{u_{in}}^{u_0} \frac {\,\textrm {d}u}{\sqrt {2h + \alpha _2 u^2 - \frac {2\gamma _2}{d+1}u^{d+1}}}= L\pmod { T} \end{equation}
$ h \lt 0$ , where
\begin{equation*} T \;:\!=\; \int _{u_{l}}^{u_{r}} \frac {1}{\sqrt {\alpha _2 u^2 - \frac {2\gamma _2}{d+1}u^{d+1} + 2h}} \,\textrm {d}u, \end{equation*}
$ u_{l}$ and
$ u_{r}$ are the intersections of the periodic orbit of the 2nd system with the
$ u$ -axis, then system (3.1) admits a pinned
$\left ( 2\left \lfloor \frac {L}{T} \right \rfloor + 1 \right )$ -hump pulse solution shown in Figure 3c .
-
-
• When
$ p_{in} \lt 0$ .
-
– If
(3.33)admits a non-degenerate root\begin{equation} \int _{u_{l}}^{u_{in}} \frac {\,\textrm {d}u}{\sqrt {2h + \alpha _2 u^2 - \frac {2\gamma _2}{d+1}u^{d+1}}} + \int _{u_{l}}^{u_0} \frac {\,\textrm {d}u}{\sqrt {2h + \alpha _2 u^2 - \frac {2\gamma _2}{d+1}u^{d+1}}}= L\pmod { T}, \end{equation}
$ h \lt 0$ , then system (3.1) admits a pinned
$\left ( \left \lfloor \frac {L}{2T} \right \rfloor + 3 \right )$ -hump pulse solution shown in Figure 3d .
-
Furthermore, to leading order, the pinned pulse
$ \big (u_{h}(\xi ), v_{h}(\xi )\big )$
is given by

where
$v_0( \xi )$
is defined by (3.4),
$u_0$
is solved from (3.9) and

where
$u_{h,\pm }(\chi )$
and
$u_{i,\pm }(\chi ,h)$
are defined in (3.25) and (3.26), respectively.
Remark 9. Theorem 4 establishes the existence of pinned pulse solutions for equation (3.1), where
$\sigma \gt 0$
and
$u_0 \gt u_{in}$
are fixed. As mentioned above, the sign of
$\sigma$
determines the direction of the fast orbit. So in a similar way, we can derive the conditions on the existence of pinned pulse solutions when
$\sigma \lt 0$
as shown in Figure 4.
Remark 10. By comparing with the autonomous case, that is,
$f(\chi )\equiv \alpha _1$
and
$g(\chi )\equiv \beta _1$
in equation (3.1), which had been studied in Veerman and Doelman [Reference Veerman and Doelman37], where

satisfy
$u_{h,\pm }(0)=u_0$
, the existence results in [Reference Veerman and Doelman37, Theorem 2.1] can be reproduced if we set
$L=0$
and
$h=0$
in Theorem 4 and Remark 9. This comparison demonstrates the validity of the results in this paper.

Figure 3. (a)–(d) Sketched plots on the different types of pinned pulses
$u_{p,0}(\chi )$
in the
$(u, p)$
plane, where
$\sigma \gt 0$
and
$u_0 \gt u_{in}$
. (a) The homoclinic orbit of the 1st system connects to the open orbit of the 2nd system defined by the energy level
$H=h$
(
$h\gt 0$
). (b) The homoclinic orbits of 1st system connects to the homoclinic orbit of the 2nd system defined by
$H=h$
(
$h=0$
). (c) The homoclinic orbits of 1st system connects to the periodic orbit of the 2nd system defined by
$H=h$
(
$h\lt 0$
) with
$p_{in}\gt 0$
. (d) The homoclinic orbits of 1st system connects to the periodic orbit the 2nd system defined by
$H=h$
(
$h\lt 0$
) with
$p_{in}\lt 0$
. (e)
$-$
(g) Singular pulse orbits
$u_{p,0}(\chi )$
in the
$(\chi , u)$
plane, in which (f)
$\lfloor \frac {L}{T} \rfloor =1$
, (g)
$\lfloor \frac {L}{2T} \rfloor =1$
.

Figure 4. (a)–(d) Sketched plots on the pinned singular homoclinic orbits
$u_{p,0}(\chi )$
in the
$(u,p)$
plane with
$\sigma \lt 0$
and
$u_0 \gt u_{in}$
. (e)–(g) Sketched plots on the pinned singular homoclinic orbits in the
$(\chi ,u)$
plane.
4. Proof of Theorem 2
In the previous section, we prove the existence of pinned pulse solutions
$(u_{h}(x), v_{h}(x))$
to system (1.2). In this section, we analyse the (in)stability of these solutions.
Linearising equation (1.2) around
$(u_{h}(x), v_{h}(x))$
yields

Then introducing the linear operator
$\mathcal{L}$
gives the associated eigenvalue problem,

In terms of the fast scale
$\xi = x/\varepsilon$
, the eigenvalue problem turns out to be

Denote
$\varphi (\xi )=(\bar {u}(\xi ),\, \bar {p}(\xi ),\, \bar {v}(\xi ),\, \bar {q}(\xi ))^{T}$
, then system (4.3) can be rewritten as

where

with

This section aims to determine the spectrum
$\sigma (\mathcal{L} )$
, which consists of the essential spectrum
$\sigma _{\textrm {ess}}(\mathcal{L} )$
and the point spectrum
$\sigma _{\textrm {pt}}(\mathcal{L} )$
. According to [Reference Kapitula and Promislow20, Theorem 3.1.11], the essential spectrum consists of all the eigenvalues such that the asymptotic matrix of (4.3) is not hyperbolic. Taking the limit
$|\xi |\rightarrow \infty$
in
$ A (\xi ,\lambda ,\varepsilon )$
, we obtain the asymptotic matrix,

Solving
$\det |\Lambda I - A_{\infty }(\lambda , \varepsilon )| = 0$
gives the eigenvalues
$\Lambda _{1, 4}(\lambda )=\pm \Lambda _{f}(\lambda )$
and
$\Lambda _{2, 3}(\lambda )=\pm \varepsilon ^2\Lambda _{s}(\lambda )$
, where

and the corresponding eigenvectors are

Thus, the essential spectrum of the eigenvalue problem (4.4) is

Under the assumptions (A1) and (A3), the essential spectrum lies entirely in the left half of the complex plane.
4.1. The Evans function
Define the complement of the essential spectrum in the complex plane
$\mathbb{C}$
by

Since
$|\Lambda _{2,3}(\lambda )|\ll |\Lambda _{1,4}(\lambda )|$
, the solutions to the eigenvalue problem (4.3) also exhibit a slow-fast nature.
Lemma 1.
For any
$\lambda \in \mathcal{C}_e$
, system (4.3) has four solutions
$ \{\varphi _{\pm ,v}(\xi ,\lambda ),\, \varphi _{\pm ,u}(\xi ,\lambda )\}$
satisfying the asymptotic conditions as follows:

Proof. Since
$A (\xi ,\lambda ,\varepsilon )\rightarrow A_{\infty }(\lambda ,\varepsilon )$
as
$\xi \rightarrow \pm \infty$
and

the conclusion follows directly from [Reference Coppel9, Theorem 4.1].
According to [Reference Kapitula and Promislow20],
$\lambda$
is an eigenvalue of
$\mathcal{L}$
if and only if there exists a non-trivial function
$\varphi \in \textrm{C}^1(\mathbb{R}, \mathbb{R}^4)$
solving system (4.4). By Lemma subsection 1, the solution
$\varphi _{+,v/u}(\xi , \lambda )$
decays exponentially to
$(0, 0, 0, 0)^T$
as
$\xi \to -\infty$
, and
$\varphi _{-,v/u}(\xi , \lambda )$
decays exponentially to
$(0, 0, 0, 0)^T$
as
$\xi \to +\infty$
. Thus,
$\{\varphi _{+,u}, \, \varphi _{+,v}\}$
and
$\{\varphi _{-,u}, \, \varphi _{-,v}\}$
span the unstable subspace
$\Phi _+(\xi , \lambda )$
and the stable subspace
$\Phi _-(\xi , \lambda )$
, respectively. The eigenfunctions of (4.4) correspond to the intersections of these subspaces. This geometric viewpoint provides the motivation for the construction of the Evans function
$E(\lambda , \varepsilon )$
, which is analytic in
$\lambda$
, and whose zeros correspond to the eigenvalues of
$\mathcal{L}$
(counted with multiplicity). These observations lead to the following definition of the Evans function:

Consequently, the problem of identifying the eigenvalues is reduced to the location the zeros of the analytic function
$E(\lambda , \varepsilon )$
(see [Reference Kapitula and Promislow20, Lemma 9.3.4]). Since
$\Lambda _{1,4}(\lambda ) = \mathcal{O}(1)$
and
$\Lambda _{2,3}(\lambda ) = \mathcal{O}(\varepsilon ^2) \ll 1$
, the functions
$\varphi _{+,v}(\xi , \lambda )$
and
$\varphi _{-,v}(\xi , \lambda )$
are uniquely determined by their asymptotic behaviour as
$\xi \to -\infty$
and
$\xi \to \infty$
.
Since
$v_{h}(\xi )$
vanishes at leading order in the slow fields, it follows that
$A (\xi ,\lambda ,\varepsilon )$
approaches a slowly varying intermediate matrix, namely,

We observe in (4.12) that the dynamics of the slow and fast variables have been separated. Furthermore, the distance between
$A_{s}(\xi , \lambda ,\varepsilon )$
and
$A_{\infty }(\lambda ,\varepsilon )$
is exponentially small.
Lemma 2. Consider

where
$\psi =(u,p,v,q)^T$
and
$A_{s}(\xi ,\lambda ,\varepsilon )$
is given in (4.12). Let
$\lambda \in \Omega$
, and then for
$\xi \lt -\frac {1}{\sqrt {\varepsilon }}$
, the stable and unstable subspaces of (4.13) are, respectively, spanned by
$\psi ^-_{+,u,v}(\xi ,\lambda )$
and
$\psi ^-_{-,u,v}(\xi ,\lambda )$
, which are given by

enjoying the following properties:

For
$\xi \gt \frac {1}{\sqrt {\varepsilon }}$
, by the reversibility symmetry, the solution spaces of (4.13) now are spanned by
$\psi ^+_{\pm ,u,v}(\xi ,\lambda )$
, where

Theorem 5. The eigenvalues are given by the zeros of the Evans function

where
$t_{f,+}(\lambda ,\varepsilon )$
is an analytic transmission function and
$t_{s,+}(\lambda ,\varepsilon )$
is meromorphic function in
$\lambda$
. These functions are defined via the asymptotic behaviour

where
$\varphi _{+,v}(\xi ,\lambda )$
has been stated in Lemma subsection 1 and
$\varphi _{+,u}(\xi ,\lambda )$
is the unique solution to (4.4), provided
$t_{f,+}(\lambda ,\varepsilon )\neq 0$
, satisfying the boundary conditions

Moreover, there exists a meromorphic transmission function
$t_{s,-}(\lambda ,\varepsilon )$
such that, to leading order,

Proof. It follows from (3.34) that there are two positive,
$\mathcal{O}(1)$
constants
$C_1$
and
$C_2$
such that

for
$\xi \in I_{s}^\pm$
. Hence, to leading order,
$\psi _{\pm ,v}(\xi , \lambda )$
and
$\psi _{\pm ,u}(\xi , \lambda )$
can be selected as basis solutions to (4.4). According to the asymptotics of
$\varphi _{+,v}(\xi ,\lambda )$
as
$\xi \rightarrow -\infty$
, we know that it remains exponentially close to
$\psi _{+,v}(\xi ,\lambda )$
for
$\xi \lt -\frac {1}{\sqrt {\varepsilon }}$
. Although the exact form of
$\varphi _{+,v}(\xi ,\lambda )$
is unknown, to leading order, it can be written as

for
$\xi \gt \frac {1}{\sqrt {\varepsilon }}$
. This proves the first result in (4.18). In order to determine
$\varphi _{+,u}(\xi )$
uniquely, it can be found that
$\varphi _{+,u}(\xi ,\lambda )$
grows obeying the speed of
$\mathcal{O}(\textrm {e}^{\Lambda _{2}(\lambda )\,\,\xi })$
. That is,
$\varphi _{+,u}(\xi ,\lambda )$
must satisfy (4.19) if
$t_{f,+}(\lambda )\neq 0$
. Hence, similar to
$\varphi _{+,v}(\xi ,\lambda )$
, we obtain (4.20) accordingly.
By Liouville’s formula and
$\textrm {Tr}\textrm {A}(x)=\Sigma _{i=1}^4 \Lambda _{i}(\lambda )$
, the Evans function turns out to be

As a consequence, the eigenvalues of (4.3) are exactly identical with the roots of
$t_{s,+}(\lambda ,\varepsilon )$
or
$t_{f,+}(\lambda ,\varepsilon )$
. Nevertheless, we will also see that some roots of
$t_{f,+}(\lambda ,\varepsilon )$
may not correspond to the true eigenvalues, as
$t_{s,+}(\lambda ,\varepsilon )$
may possess poles at these values, which cancel the zeros.
4.2.
The fast transmission function
$t_{f,+}(\lambda ,\varepsilon )$
In this section, we determine the zeros of the fast transmission function
$t_{f,+}(\lambda ,\varepsilon ).$
Let
$\varepsilon \rightarrow 0$
in system (4.3), we get

which is a second-order inhomogeneous differential equation. Here, we have used the fact that
$u_h(\xi )=u_0$
and
$v_h(\xi )=v_{0}(\xi )$
for
$\xi \in I_f$
as
$\varepsilon \rightarrow 0$
. Equation (4.21) corresponds to the fast limiting eigenvalue problem in the singular limit of system (4.3), which is a classical Sturm–Liouville eigenvalue problem. Thus, by Sturm–Liouville theory, the eigenvalues of
$\mathcal{L}_f$
can be enumerated in a strictly descending order,

Moreover, the eigenfunction
$\bar {v}_i(\xi )$
corresponding to
$\lambda ^f_{i}$
has exactly
$i$
distinct zeroes and is even or odd depending on whether
$i$
is even or odd, respectively.
The homogeneous problem associated with system (4.21) can be written as

where
$\phi =(v,q)^T$
and
$A_{f}(\xi , \lambda )$
is a
$2\times 2$
matrix, which is exactly the lower
$2\times 2$
block of
$A(\xi , \lambda ,\varepsilon )$
in the limit
$\varepsilon \rightarrow 0.$
Obviously, system (4.22) has two solutions
$\phi _{+}(\xi , \lambda )$
,
$\phi _{-}(\xi , \lambda )$
satisfying

and there exists an analytic function
$t_{f}(\lambda )$
such that

The Evans function associated with this problem is given by

By construction, the leading order behaviour of
$t_{f,+}(\lambda ,\varepsilon )$
is determined by
$t_{f}(\lambda )$
. It thus follows from (4.25) that
$t_{f}(\lambda )=0$
if and only if
$E_{f}(\lambda )=0$
. Hence, the eigenvalues of (4.22) correspond to the zeros of
$t_{f}(\lambda )=0$
. That is, the roots of the function
$t_{f}(\lambda )$
are given by
$\lambda ^f_{i}$
,
$i=0,1,\ldots ,N$
.
Lemma 3.
Let
$0\lt \varepsilon \ll 1$
, then the fast transmission function is given by

with
$\tilde {t}_f(\lambda ,\varepsilon )\neq 0$
, and
$\lambda ^f_{i}(\varepsilon )$
has the following regular expansion:

for
$i=0,1,\ldots ,N.$
The proof of this lemma is analogous to those in [Reference Doelman, Gardner and Kaper12–Reference Doelman, Iron and Nishiura14] by using the ‘elephant trunk’ procedure [Reference Alexander, Gardner and Jones1, Reference Gardner and Jones18]. It is important to note that
$\lambda ^f_{1}(\varepsilon )\equiv 0$
is no longer valid since system (1.2) now is non-autonomous, leading to the loss of translational invariance. As a result, under the perturbation, the eigenvalue
$\lambda ^f_{1}(\varepsilon )$
may be shifted to the one having the positive real part. Under certain situations, standard perturbation methods such as the Lin–Sandstede method [Reference Bastiaansen, Chirilus-Bruckner and Doelman3, Reference Kapitula and Promislow20, Reference Li and Yu23] can be applied to estimate the location of these eigenvalues. However, due to the piecewise-smooth nature of system (1.2), no effective methods currently exist for precisely determining the location of these eigenvalues.
Remark 11. While the Lin–Sandstede method can be used to detect the eigenvalue bifurcating from
$\lambda = 0$
under perturbations, its validity depends heavily on the assumption that the eigenfunction associated with the perturbed eigenvalue remains a small deformation of the unperturbed one. This assumption enables a regular perturbation approach, where a solvability condition – typically of Fredholm type – is imposed to determine the first-order correction
$\lambda =\lambda _1^f(\varepsilon )$
. However, in the singularly perturbed system studied here, this assumption does not hold. As shown in the previous section, the pulse solution undergoes a transition when it passes the interfaces at
$\chi =\pm L$
. Consequently, the eigenfunction associated with
$\lambda = \lambda _1^f(\varepsilon )$
does not converge to the eigenfunction associated with
$\lambda = 0$
as
$\varepsilon \to 0$
. So Lin–Sandstede method is not applicable here. We adopt an alternative technique to study the spectrum of the linearised operator.
Although
$\lambda ^f_0(\varepsilon )$
is a simple zero of the function
$t_{f,+}(\lambda , \varepsilon )$
, this does not imply that the function
$E(\lambda , \varepsilon )$
also vanishes at this point. This is due to the fact that
$\lambda = \lambda ^f_0(\varepsilon )$
is a first-order pole of the transmission function
$t_{s,+}(\lambda )$
and therefore cannot be an eigenvalue of the full system. This phenomenon is known as the ‘NLEP Paradox’ in literature.
According to (4.23)–(4.24), we know that
$\phi _{+}(\xi ,\lambda )$
and
$\phi _{-}(\xi ,\lambda )$
form a fundamental set of linearly independent solutions to the homogeneous system (4.22) for
$\lambda \neq \lambda ^f_{i}$
,
$i=0,1,\ldots N$
. Consequently, by employing the method of constant variation we can derive a bounded solution to the heterogeneous problem (4.21).
Lemma 4.
For
$\lambda \neq \lambda ^f_{i}$
,
$i=0,1,\ldots N$
, the unique bounded solution
$v_{in}(\xi ,\lambda )$
to the inhomogeneous problem (4.21) is given by

where
$v_{\pm }(\xi , \lambda )$
is the
$v$
-component of
$\phi _{\pm }(\xi , \lambda )=(v_{\pm }(\xi , \lambda ),q_{\pm }(\xi , \lambda ))$
, and
$E_f(\lambda )$
is defined as in (4.25).
Lemma 5.
(See [Reference de Rijk, Doelman and Rademacher10]) The bounded solution
$v_{in}(\xi , \cdot )$
is meromorphic in the region
$\mathcal{C}_e$
for fixed
$\xi \in \mathbb{R}$
and analytic on
$\mathcal{C}_e\backslash \{\lambda ^f_{i}\}_{i=0}^N$
for
$\xi \in \mathbb{R}$
. Furthermore, when
$i$
is even,
$\lambda =\lambda ^f_{i}$
is a pole of
$v_{in}(\xi , \lambda )$
, while when
$i$
is odd, it is a removable singularity.
Proof. Since
$\mathcal{L}_f - \lambda$
is a Fredholm operator of index zero for
$\lambda \in \mathcal{C}_e$
, and
$\mathcal{L}_f - \lambda$
is invertible if and only if
$\lambda \in \mathcal{C}_e\backslash \{\lambda ^f_{i}\}_{i=0}^N$
, it follows that
$(\mathcal{L}_f - \lambda )^{-1}$
is analytic on
$\mathcal{C}_e\backslash \{\lambda ^f_{i}\}_{i=0}^N$
and meromorphic on the entire region
$\mathcal{C}_e$
. Since
$\mathcal{L}_f - \lambda$
is a self-adjoint operator, it follows from the Fredholm alternative [Reference Kapitula and Promislow20, Theorem 2.2.1] that a bounded solution to the inhomogeneous equation exists at
$\lambda \in \{\lambda ^f_{i}\}_{i=0}^N$
if and only if the following solvability condition is satisfied:

where
$\bar {v}_{i}(\xi , \lambda )\in \ker (\mathcal{L}_f-\lambda )$
. According to Sturm–Liouville theory, if
$i$
is even, the corresponding eigenfunction
$\bar {v}_i(\xi )$
is an even function, and if
$i$
is odd, it is an odd function. Since the homoclinic solution
$v_0(\xi )$
is even, the solvability condition does not hold when
$i$
is even. Therefore,
$\lambda = \lambda ^f_{i}$
is a pole of
$v_{in}(\xi , \lambda )$
. When
$i$
is odd, the solvability condition holds, and
$\lambda =\lambda ^f_{i}$
is a removable singularity.
Remark 12. The bounded solution
$v_{in}(\xi ,\lambda )$
to the inhomogeneous problem (4.21) forms one of the key ingredients of the slow transmission function
$t_{s,+}(\lambda ,\varepsilon )$
(see the subsequent subsection).
4.3.
The slow transmission function
$t_{s,+}(\lambda ,\varepsilon )$
We will see that the slow transmission function
$t_{s,+}(\lambda ,\varepsilon )$
can be determined by matching the slow and fast orbits, respectively, of the following slow limit eigenvalue problem:

and the fast limit eigenvalue problem (4.21).
Recall that, on the slow intervals
$I_s^{\pm }$
,

To simplify the notation and to enhance the clarity of formulation, we utilise the variable
$\chi =\varepsilon ^2 \xi$
and define
$ (u_{\pm }(\chi , \lambda ), p_{\pm }(\chi , \lambda ))$
.
Denote

with
$p_{\pm }(\chi , \lambda )=\frac{\text{d}}{\text{d}\chi }u_{\pm }(\chi , \lambda )$
; thus,
$\left (u_{\pm }(\chi , \lambda ), p_{\pm }(\chi , \lambda )\right )$
is the unique solution to the slow limit eigenvalue problem (4.29) satisfying

Lemma 6.
Define
$u_{\pm }(0, \lambda )=u_{\pm }(\chi =0)$
and
$p_{\pm }(0, \lambda )=p_{\pm }(\chi =0)$
, then the leading order of the slow transmission function can be given by

where

Proof. It follows from Theorem 5 that there exists a positive constant
$K$
independent of
$\varepsilon$
such that

for
$\xi \in I_s^-$
, and

for
$\xi \in I_s^+$
. Accordingly, we can determine the explicit representation of
$t_{s,+}(\lambda ,\varepsilon )$
by matching the fast and slow orbits. More precisely, to compute
$t_{s,+}(\lambda , \varepsilon )$
, we need to track the changes of
$\bar {u}_{+}(\xi , \lambda )$
and
$\bar {p}_{+}(\xi , \lambda )$
during the fast transition, that is,
$\xi \in I_f$
.
We define the slow difference function
$\Delta _s$
by

First, it follows from (4.4) that
$\bar {u}_+(\xi , \lambda )$
remains constant to leading order during the fast field
$I_f$
. As a consequence,

This together with

gives

Thus, the first matching condition (4.33) can be written as

So the first relation between
$t_{s,+}(\lambda )$
and
$t_{s,-}(\lambda )$
can be given by

On the other hand, the derivative
$\bar {p}_{+}(\xi , \lambda )$
changes rapidly in the fast field
$I_f$
. It can be determined through the following two ways. First, we use the information on
$\varphi _{+,u}(\xi , \lambda )$
in the slow field to compute it. The change
$\Delta _{s}\bar {p}$
, to leading order, can be approximated by the difference of
$\bar {p}_{+}(\xi , \lambda )$
between two ends of the slow field, that is,

Second, the accumulated jump concerning the first-order derivative can be obtained by integrating
$\overline {u}_{\xi \xi }$
over the fast field, that is, to leading order,

By combining (4.38) and (4.39), one can get (4.30) up to the leading order.
Remark 13. It is worthy to note that only the roots of
$t_{s,+}(\lambda ,0)$
are not sufficient to decide the spectral stability of the pinned pulse of (1.2) due to the fact that the sign of
$\lambda ^f_{1}(\varepsilon )$
cannot be determined.
Remark 14. If
$F_v(u,v)\equiv 0$
for all
$u,\,v\gt 0$
, then

which implies that the transmission function
$t_{s,+}(\lambda ,\varepsilon )$
is only determined by the slow reduced eigenvalues problem (4.29). As a consequence,
$t_{s,+}(\lambda ,\varepsilon )$
is analytic on
$\mathcal{C}_e$
, and the zeros of
$t_{f,+}(\lambda ,\varepsilon )$
cannot be cancelled by the poles of
$t_{s,+}(\lambda ,\varepsilon )$
. In this case, we conclude that the linearised eigenvalue problem can be completely separated into the slow and fast reduced eigenvalue problems without the interaction between them. Thus, the zeros of
$t_{f,+}(\lambda ,\varepsilon )$
with positive real parts yield the spectral instability directly.
5. (In)stability of pinned pulses in non-autonomous GM equation
Consider

whose existence has been set up in Section 3. The associated eigenvalue problem is

where
$\varphi (\xi )=(\bar {u}(\xi ), \bar {p}(\xi ), \bar {v}(\xi ), \bar {q}(\xi ))^{T}$
, and

By (5.2), the essential spectrum turns out to be

Now (4.21) becomes

Obviously, the eigenvalues associated with the operator
$\mathcal{L}_f$
are well-known, namely,
$\lambda ^f_{0}=5/4$
,
$\lambda ^f_{1}=0$
and
$\lambda ^f_{2}=-3/4$
.
As treated in [Reference Veerman and Doelman37], we can obtain the following result.
Lemma 7.
For
$0\lt \varepsilon \ll 1$
, the fast transmission function can be given by

with
$\tilde {t}_f(\lambda ,\varepsilon )\neq 0$
, and
$\lambda ^f_{i}(\varepsilon )$
has the following regular expansion:

In particular, for
$\lambda \neq \lambda ^f_{i}, i=0,1,2$
, the unique solution of
$v_{in}(\xi ,\lambda )$
to the heterogeneous problem (4.21) now is

where
$v_{\pm }(\zeta , \lambda )=c_{\pm }(\lambda )P_3^{-2\sqrt {1+\lambda }}\left (\pm \zeta \right )$
,
$\zeta =\tanh \frac {\xi }{2}$
,
$P_3^{-2\sqrt {1+\lambda }}\left (\pm \zeta \right )$
are the Legendre functions and
$c_{+}(\lambda )c_{-}(\lambda )=-\frac {1}{2}\Gamma (4+2\sqrt {1+\lambda })\Gamma (\!-3+2\sqrt {1+\lambda })$
.
Lemma 8.
Concerning (5.1), the integral
$\mathcal{G}(\lambda )$
defined in (4.31) now is

with

By Lemma 2, we denote
$u_{\pm }(\chi ,\lambda )$
the solutions to the following equation:

with the boundary condition
$\lim _{\chi \rightarrow -\infty }u_{\pm }(\chi )\textrm {e}^{\mp \sqrt {\alpha _1+\lambda }\chi }=1$
. It then follows from Theorem 2 that
$t_{s,+}(\lambda )=0$
if and only if
$u_+(0,\lambda )=0$
or

Theorem 6. Let the conditions in Theorem 3 be satisfied, then
-
(i) When
$u_{0}(h)=u^+_{0}(h)$ , if
(5.12)admits\begin{equation} \begin{aligned} \sqrt {\frac {\alpha _2+\lambda }{\alpha _2+\sqrt {\alpha _2^2+72\sigma ^2h}}}\left (1-\frac {2} {\left (1+\frac {2}{\sqrt {\frac {\alpha _2+\lambda }{\alpha _1+\lambda }}-1}\right )\textrm { e}^{2\sqrt {\alpha _2+\lambda }}+1}\right ) = 9 \sqrt {2} \mathcal{R}(\lambda ) \end{aligned} \end{equation}
$\lambda \in \{\lambda \mid \text{Re}\, \lambda \gt 0\}$ , then the pinned pulse in equation (5.1) is unstable.
-
(ii) When
$u_{0}(h)=u^-_{0}(h)$ , if
(5.13)admits\begin{equation} \begin{aligned} \sqrt {\frac {\alpha _2 + \lambda }{\alpha _2 - \sqrt {\alpha _2^2 + 72 \sigma ^2 h}}} \left ( 1 - \frac {2}{\left (1 + \frac {2}{\sqrt {\frac {\alpha _2 + \lambda }{\alpha _1 + \lambda }} - 1}\right ) e^{2 \sqrt {\alpha _2 + \lambda }} + 1}\right ) = 9 \sqrt {2} \mathcal{R}(\lambda ) \end{aligned} \end{equation}
$\lambda \in \{\lambda \mid \text{Re}\, \lambda \gt 0\}$ , then the pinned pulse in equation (5.1) is unstable.
Proof. Now equation (5.11) turns out to be

Hence,

and

We choose the free parameters
$M_i, N_i, i=1, 2$
such that
$u_{\pm }(\chi , \lambda )$
is continuously differentiable at
$\chi =-L$
. Consequently,

and further by the reversibility symmetry, we get

This results in
$u_+(0, \lambda )=M_1+N_1\neq 0$
and

Combining (3.12) with (5.9) gives

Based on the results of Theorem 2, we accordingly get (5.12) and (5.13).
Theorem 7.
Let
$\varepsilon \gt 0$
be sufficiently small, and let
$\alpha _1$
,
$\alpha _2\gt 0$
,
$\gamma _1,\,L\gt 0$
,
$\gamma _2=0$
and
$d\gt 1$
be fixed such that equation (5.1) admits a pinned pulse solution, then this pulse is unstable if

or

possesses
$\lambda \in \{\lambda \mid \text{Re}\,\lambda \gt 0 \}$
, where

in which
$P_\nu ^\mu (z)$
and
$P_\nu ^\mu (\!-z)$
are the Legendre functions and
$x_*$
is defined by (3.25).
Proof. In the slow interval
$(\!-\infty ,-L)$
, to leading order, equation (5.11) is governed by

whose solution is denoted by
$u_{h,-}(\chi )$
,
Introducing the Hopf-cole transformation

to equation (5.22) yields the Legendre differential equation,

where

The solutions of (5.24) can be given in terms of the Legendre functions
$P_\nu ^\mu (z)$
and
$P_\nu ^\mu (\!-z)$
. In light of the results in Lemma 2, we have

satisfying
$u_+(\chi )\rightarrow 0$
as
$\chi \rightarrow -\infty .$
To further simplify the notation, let us introduce

in which

Moreover, by following Lemma 6, we have

where

As a consequence,

This means that
$u_+(0,\lambda )p_-(0,\lambda )-u_-(0,\lambda )p_+(0,\lambda )=2\sqrt {\alpha _2+\lambda }(N_3^2-M_3^2)\neq 0$
and

Define two functions
$\textrm{B}_{\pm }(\chi )$
and their derivatives
$\textrm{B}^{\prime }_{\pm }(\chi )$
, respectively, by

Theorem 8.
Let the conditions in Theorem
4
be fulfilled. If the length of the heterogeneity
$L$
satisfies
$h(L)=\mathcal{O}(\varepsilon )$
, then the pinned pulse solution of equation (5.1) is unstable if
$\lambda \in \{\lambda \,|\,\text{Re}\,\lambda \gt 0 \,\}$
,

or

admits a root
$\lambda$
with positive real part, where
$u(\!-L, \lambda )$
and
$p(\!-L, \lambda )\,$
are defined in equation (5.27), and
$y_0=\tanh \left (\frac {(d-1)\sqrt {\alpha _2}}{2}y_*\right )$
. Furthermore, the functions
$\textrm{B}_1(\lambda ),$
$i=1,2,3,4$
, are given by

Proof. First, let us consider the simple case, namely,
$h(L)=0$
. In this case, equation (5.11) can be written as

where the solutions
$u_{i,-}(\chi ,0 )$
and
$u_{h,-}(\chi )$
are defined in equation (3.25) and (3.27), respectively. According to the result of Lemma 7, the solution
$u_{+}(\xi , \lambda )$
is given by

where

and

with
$u(\!-L, \lambda )$
and
$p(\!-L, \lambda )$
are given by (5.27). Note that the expression for
$ u_{+}(\chi , \lambda )$
in equation (5.36) is valid only when
$ \lambda \neq \lambda _j$
, for
$ j = 0, 1, \ldots , J$
, where
$ J \lt \nu \leq J + 1$
and
$ \lambda _j = \frac {1}{4} \left [ (d+1) - j(d-1) \right ]^2 - 1$
. This restriction is generated since, when
$ \lambda = \lambda _j$
, the pair of functions
$ \{ P_{\nu }^{\mu _2}(z), P_{\nu }^{\mu _2}(\!-z) \}$
is insufficient to span the solution space of equation (5.35). Thus, we have

and

It follows that

We next consider the general case, that is,
$h(L) = \mathcal{O}(\varepsilon )$
. We can further derive the solution to equation (5.11) by using perturbation analysis. First, recall that the Hamiltonian value
$h(L) = 0$
corresponds to the homoclinic orbit in the 2nd system. Thus, for
$h(L) = \mathcal{O}(\varepsilon )$
, the orbit is located near this homoclinic orbit. Moreover, when
$\varepsilon \gt 0$
is sufficiently small, the distance between the two orbits becomes very small. Specifically, when
$h(L) = \mathcal{O}(\varepsilon )$
, equation (5.11) can be regarded as a small perturbation of equation (5.35). It is easy to verify that equation (5.35) exhibits exponential dichotomy on the regions
$\mathbb{R}_{\pm }$
. By the principle of exponential dichotomy and its robustness, the stable and unstable spaces of equation (5.11) are
$\mathcal{O}(\varepsilon )$
-close to those of equation (5.35). Thus, equations (5.39) and (5.40) remain valid when
$h(L) = \mathcal{O}(\varepsilon )$
.
Let

We next use
$y_0$
to express
$u_0$
. Using the fact that the jump point
$(u_0, p_0)$
is the intersection of the Take-off curve and the orbit defined by
$H_s(u, p) = 0$
, we obtain the following relations:

Combining (3.27) and (5.37) gives
$y^2(\chi )=\frac {1}{\alpha _2}\frac {u^2_{\chi }}{u^2}$
. Therefore, we have the following relationship:

which gives

By combining the conclusions from Theorem 2 and Lemma 8, we obtain equations (5.33) and (5.34). Based on (5.33) or (5.34), we can identify the signs of the eigenvalues
$\lambda$
numerically under certain parameter values. In this manner, we can judge whether the pinned pulses is stable or unstable under the present parameter condition.
Remark 15. If the non-autonomous GM equation reduces to the autonomous one, that is,
$f(\chi )\equiv \alpha _1$
and
$g(\chi )\equiv \beta _1$
in equation (5.1), which had been studied in Veerman and Doelman [Reference Veerman and Doelman37], in this case, we have
$L=0$
. Hence, (5.39) and (5.40), respectively, reduce to

with
$z_*=\textrm {sign}(\sigma )\tanh \left (\frac {1}{2}(d-1)\sqrt {\alpha _1}\, x_* \right )$
. That is, the instability criterion (5.33) and (5.34), respectively, reduce to

or

which are identical with which in [Reference Veerman and Doelman37, Corollary 4.1] for the autonomous GM equation with a slow nonlinearity.
Financial support
The research is supported by the National Science Foundation (NSF) of China (No. 12271096), the NSF of Fujian Province (No. 2022J02028) and the Young Top Talent of Fu-jian Young Eagle Program.
Competing interests
The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.