Abstract

In some pattern-forming systems, for some parameter values, patterns form with two wavelengths, while for other parameter values, there is only one wavelength. The transition between these can be organized by a codimension-three point at which the marginal stability curve has a quartic minimum. We develop a model equation to explore this situation, based on the Swift–Hohenberg equation; the model contains, amongst other things, snaking branches of patterns of one wavelength localized in a background of patterns of another wavelength. In the small-amplitude limit, the amplitude equation for the model is a generalized Ginzburg–Landau equation with fourth-order spatial derivatives, which can take the form of a complex Swift–Hohenberg equation with real coefficients. Localized solutions in this amplitude equation help interpret the localized patterns in the model. This work extends recent efforts to investigate snaking behaviour in pattern-forming systems where two different stable non-trivial patterns exist at the same parameter values.

1. Introduction

Pattern formation most commonly occurs with a single wavelength, as in for example zebra stripes, Rayleigh–Bénard convection and the Taylor–Couette flow (Hoyle, 2006). In these examples, there is a featureless basic state that loses stability to waves with a non-zero wavelength as a control parameter is increased. Typically the marginal stability curve, which separates stable from unstable waves depending on their wavelength and the control parameter, has a quadratic minimum.

In recent years, it has been recognized that pattern formation with two length scales can lead to a wide variety of complex and interesting patterns, such as superlattice patterns, quasipatterns and spatio-temporal chaos (see for example Castelino et al., 2020, and references therein). Having two length scales can arise in different ways: in the Faraday wave problem with multi-frequency forcing, for example, patterns with the two length scales arise in response to different components of the forcing (Edwards & Fauve, 1994; Rucklidge & Silber, 2009; Skeldon & Rucklidge, 2015; Topaz & Silber, 2002). Another possibility is that the quadratic minimum in the marginal stability curve can change to a quadratic maximum with two nearby quadratic minima at the two length scales. This transition can occur via a quartic minimum, and is found in the magnetized Taylor–Couette experiement (Mamatsashvili et al., 2019; Stefani et al., 2009), Lapwood–Prats convection (Rees & Mojtabi, 2013), binary phase field crystals (Holl et al., 2021), surface waves in ferrofluids (Raitt & Riecke, 1997) and nonlinear optics (Kozyreff et al., 2009). This paper is concerned first with developing and analyzing a model that contains this transition in as simple a form as possible, and second with investigating localized patterns in the model.

Problems with a single length scale where the pattern-forming bifurcation is subcritical can have parameter intervals where both the featureless and the patterned solutions are stable. In this case, it is possible to find localized solutions consisting of a region of a spatially periodic pattern embedded in a spatially homogeneous background (see Dawes, 2010; Knobloch, 2015, for reviews). With two length scales, there is a wider variety of possibilities, including having patterns with one wavelength embedded in a pattern with a different wavelength. This phenomenon has been observed in Rayleigh–Bénard convection in a long, thin channel or slot (Hegseth et al., 1992), in large-aspect-ratio thermosolutal convection (Spina et al., 1998), and has been explored in the context of generalized Ginzburg–Landau models (Bortolozzo et al., 2006; Kozyreff et al., 2009; Raitt & Riecke, 1995, 1997; Riecke, 1990).

A useful mathematical tool for studying pattern formation is the construction of model equations that display qualitatively similar behaviour as the physical system under consideration, but whose analysis is more tractable. Perhaps the most ubiquitous of such model equations is the Swift–Hohenberg (SH) equation (Swift & Hohenberg, 1977), originally introduced as a model of thermal fluctuations near the onset of Rayleigh–Bénard convection. It has been used extensively in the study of localized patterns, starting with the work of Hilali et al. (1995), Crawford & Riecke (1999) and Woods & Champneys (1999). The equation (in one dimension) is
(1.1)
where u(x,t)R is an order parameter that represents the pattern, μ is the driving parameter and n2 and n3 are parameters controlling the nonlinear terms (typically n3=1). We consider equation (1.1) subject to periodic boundary conditions on a domain x[0,L].
The featureless (or trivial) solution u=0 is stable for μ<0. Small amplitude perturbations of the form eσt+ikx grow as σ=μ(1k2)2, so for μ>0 the maximum growth rate is at critical wavenumber k=1 (independent of μ), and if μ>0, a range of wavenumbers will grow exponentially until nonlinear effects become important. The marginal stability curve is found by setting σ=0, so
which has a quadratic minimum at k=1. At μ=0 the trivial solution undergoes a pitchfork bifurcation, creating a branch of spatially periodic solutions, which is stable if it is supercritical, and unstable if not.
In large domains (L1), and with small amplitude solutions (u=O(ϵ)), standard weakly non-linear theory can be applied. The pattern is written, with scaled space, time and parameter, as
(1.2)
where c.c. refers to the complex conjugate and h.o.t. refers to higher-order terms. In this limit, the solvability condition for A at third-order in ϵ results in the Ginzburg–Landau (GL) equation (Cross & Hohenberg, 1993):
(1.3)
where subscripts T and X refer to partial derivatives. This equation governs the long-wavelength slow evolution of the amplitude of solutions of (1.1). We define the coefficient of the nonlinear term
which determines the criticality of the bifurcation at μ=0: if nSH<0 the bifurcation is supercritical, and if nSH>0, the bifurcation is subcritical. In the supercritical case, the GL equation gives nonlinear stability of striped patterns to long-wavelength perturbations (the Eckhaus instability, see Eckhaus, 1965). In the subcritical case, the GL equation allows localized sech-profile solutions to equation (1.1) that can be continued in μ, leading to the well known homoclinic “snaking” structure of localized solutions of the Swift–Hohenberg and many other pattern-forming problems (Beck et al., 2009; Burke & Knobloch, 2006, 2007b; Chapman & Kozyreff, 2009; Coullet et al., 2000; Kozyreff & Chapman, 2006; Woods & Champneys, 1999).
The GL equation (1.3) is one standard tool useful in the analysis of the SH equation (1.1). There are three others that we mention briefly. First, the SH equation as written above is variational in time, and admits a Lyapunov functional:
(1.4)
With n3<0, it can easily be shown that F[u] is bounded below and that it is a decreasing function of time:
Equilibrium states correspond to stationary points of F, and those coinciding with local minima of F must necessarily be stable. A front connecting two patterns with different values of F will tend to move towards (and so eliminate) the pattern with the larger value. Localized solutions are found near the Maxwell point, where the pattern and the zero state have the same value of F, and near this point the difference in F is small enough that the front becomes pinned (Pomeau, 1986) to the underlying pattern.
The second tool is the observation that the steady Swift–Hohenberg equation (1.1) admits a first integral in space. Multiplying the time-independent version of (1.1) through by ux, and integrating with respect to x, yields
and so dHdx=0. The quantity H is sometimes referred to as the Hamiltonian for the steady version of (1.1), since there is a change of coordinates under which the system has Hamiltonian structure. If there is a steady front connecting two patterns, the condition dHdx=0 means that the two patterns must have the same value of H.

The third useful tool is to note, again for the time-independent version of (1.1), that there is a Hamiltonian–Hopf bifurcation in space as μ crosses zero. At the bifurcation point, there is a pair of double spatial eigenvalues ±i, and the normal form can be written as a pair of first-order ODEs in x for two complex variables. A bifurcation analysis performed by Iooss & Pérouème (1993) and Iooss & Adelmeyer (1998), and extended by Woods & Champneys (1999), provides a geometrical interpretation of the solutions of the normal form. In particular, there are parameter values where there are solutions of the SH equation that are homoclinic to the origin as x± (Burke & Knobloch, 2007a): these homoclinic orbits represent localized solutions.

The existence and bifurcation structure of localized solutions in the Swift–Hohenberg equation is now well understood: see Dawes (2010) and Knobloch (2015) for reviews. In this paper, we consider a generalized version of the Swift–Hohenberg equation that allows a quartic minimum of the marginal stability curve (§2). Unfolding this quartic minimum, and using tools such as generalized versions of the Ginzburg–Landau equation (§3), the Lyapunov function and the first integral introduced above (§4), allows us to identify parameter regimes where we can find patterns of one wavenumber localized in a background of patterns with a different wavenumber (§5). We compare our work to a derivation of the amplitude equation for a related problem, the Lugiato–Lefever equation, in §6. We discuss the significance of our results in §7. We include normal form calculations in Appendix  A, but we have not found first integrals of the normal form, and so we have not been able to put it to immediate use.

2. The model equation

In this section, we build a model equation to explore the unfolding of the quartic minimum in the marginal stability curve. We start with the SH equation (1.1), modified to allow a marginal stability that can change from having one to having two minima, and then add a selection of nonlinear terms.

2.1 Linear terms

As a starting point, we consider a linear part of the PDE based on the polynomial
where K=k2. This is the simplest polynomial that has quartic minima at k=±1. In the model equation, 1K will become 1+xx. The quartic minimum at K=1 can be unfolded by adding two small terms to the equation, yielding
(2.1)
with f1=f2=0 at the quartic minimum. In principle, small terms f0, f3K3 and f4K4 could be added as well, but f0 can be absorbed into the bifurcation parameter μ, f3 can be eliminated by making a small shift in K by 14f3, and f4 can be absorbed by an overall scaling.
Before writing down the PDE model, we consider the transition from having one minimum to two. The condition that p(K) has two minima with a maximum in between is the same as the condition that the derivative p(K) has three distinct real roots. Now
and the condition that a cubic polynomial has three distinct real roots is that its discriminant should be positive. The boundary, where the discriminant is zero, occurs where
In addition, when the discriminant is positive, p(K) has two minima at K1 and K3 with a maximum at K2 in between, with K1<K2<K3, K1+K2+K3=3 and K1K2K3=114f1 (found from the relationship between the roots and coefficients of the cubic p(K)=0). Manipulating the conditions p(K1)=0, p(K3)=0 and p(K1)=p(K3) leads to the conclusion that the two minima are equal when
that is, the intermediate maximum is at K=1 and the two minima are equally spaced on each side.
The next step is to convert the polynomial to the linear operator of the model equation:
A mode u=eσt+ikx has growth rate σ=μ(1k2)4f1k2f2k4=μp(k2), connecting the dispersion relation to the polynomial p(K) discussed above. Marginal stability, when σ=0, occurs when μ=p(k2), and Figure 1 shows examples of the marginal stability curves in the (f1+2f2,f2) parameter plane. The discriminant is positive, and there is a double minimum, within the cusp-shaped region below the solid curves, and the two minima are equal on the dashed line. The cusp, the point at which there is a quartic minimum with f1=f2=0, represents a codimension-three bifurcation as μ crosses zero.
Five examples of the marginal stability curves $\mu =p(K)$ in different regions of the $(f_{1}+2f_{2},f_{2})$ parameter space. The curves (blue in black boxes) shown are: $(i)$ single minimum, $(ii)$ quartic minimum, $(iii)$ double minima, with the left minimum being the lower, $(iv)$ double minima, with both minima at the same height, and $(v)$ a single minimum and an inflexion point. The specific parameters for each example are shown as blue dots. The solid lines indicate where the discriminant is zero (the derivative has a double zero), and the dashed line ($f_{1}+2f_{2}=0$, $f_2<0$) indicates where the two minima exist and have the same height.
Fig. 1.

Five examples of the marginal stability curves μ=p(K) in different regions of the (f1+2f2,f2) parameter space. The curves (blue in black boxes) shown are: (i) single minimum, (ii) quartic minimum, (iii) double minima, with the left minimum being the lower, (iv) double minima, with both minima at the same height, and (v) a single minimum and an inflexion point. The specific parameters for each example are shown as blue dots. The solid lines indicate where the discriminant is zero (the derivative has a double zero), and the dashed line (f1+2f2=0, f2<0) indicates where the two minima exist and have the same height.

However, a model based solely on this would have a wavenumber for maximum growth rate for solutions that did not depend on μ. Regions of secondary instability of patterns are organized around the curve of maximum growth rate, and it is important that this curve is modelled correctly. In pattern-forming problems with a quadratic minimum, the wavenumber for maximum growth rate typically depends only linearly on μ, so with μ1, this lack of dependence on μ in the SH equation is reasonable. However, with a quartic minimum, Proctor (1991) argues that additional terms should be included in the unfolding of a quartic minimum in order to allow the wavenumber of maximum growth rate to depend on μ. For this reason, we propose the modified linear operator
(2.2)
where we will call the extra term, proportional to μp, the Proctor term, although the form of this extra term differs from that proposed by Proctor (1991). The dispersion relation is now
and so
The curve of maximum (or minimum) growth rate is then defined by dσdk2=0, i.e.,
For μp0 and f1=f2=0 (Figure 2a), this curve of maximum growth rate is a cubic and so is tangent to the neutral stability curve, while for non-zero f1 and f2 (Figure 2b), the curves of maximum growth rate intersect the two minima in the marginal stability curve transversally.
Marginal stability curve (solid) and curve of maximum (or minimum) growth (dashed) for (2.2), for (a) $f_{1} = f_{2} = 0$, and (b) $f_{1}=0.05$ and $f_{2}=-0.0235$. The coefficient $\mu _{p} = -0.1$ in both cases. Two of the three dashed lines in (b) join together above the top of the frame.
Fig. 2.

Marginal stability curve (solid) and curve of maximum (or minimum) growth (dashed) for (2.2), for (a) f1=f2=0, and (b) f1=0.05 and f2=0.0235. The coefficient μp=0.1 in both cases. Two of the three dashed lines in (b) join together above the top of the frame.

Including the Proctor term influences the shape of the marginal stability curve. Throughout most of this paper we will assume that μp is small enough not to influence the behaviour of solutions, apart from the asymptotic analysis in §3.3 (where the Proctor term necessarily appears as a higher order term), and a numerical consideration in §5.2.

The linear part of the model derived so far is symmetric under spatial reversals, i.e., x and x are equivalent. Some systems break reflection symmetry, for example, the Taylor–Couette system in the presence of an azimuthal and axial magnetic field (Mamatsashvili et al., 2019; Stefani et al., 2009); this can be modelled by including terms such as (fd0fd1xx)ux, leading to drifting solutions. However, the reflection symmetry is useful in the analysis of the model equation, so we do not include drift terms here.

2.2 Nonlinear terms

Nonlinear terms in the model will saturate exponentially growing solutions at finite amplitude, and should respect the symmetry (or lack thereof) of any underlying physical systems. The usual SH non-linear term is u3, allowing uu symmetry, but here we break this symmetry and chose
as nonlinear terms for the model. There are many more possibilities for nonlinear terms, both with and without variational structure, that have been used by many authors (Burke & Dawes, 2012; Crawford & Riecke, 1999; Knobloch, 1990; Rucklidge et al., 2012) for similar equations in different contexts. For the purposes of this paper, we consider only the nonlinear terms above, retaining the coefficients n2 and n3 as parameters. We focus on the case where the bifurcation is supercritical, which limits the range of values that n2 and n3 can take. In particular, we must have
(2.3)
derived in §3.1. For numerical examples, we will take n2=0.1 and n3=1 throughout.
We now have our complete model equation, written here with the Proctor term
(2.4)
and without
(2.5)
For the remainder of this paper, we will concentrate mostly on the parameter values where the marginal stability curves have two minima and there is bistability between patterns of different wavelength, leading to the possibility of these patterns coexisting in separate parts of the domain. Although these equations are related to the two length scale models of Lifshitz & Petrich (1997) and Rucklidge et al. (2012), our model cannot be derived from these simply by setting the two length scales to be equal.

3. Weakly nonlinear analysis

In this section, we compute weakly nonlinear solutions for the model (2.5) by deriving a generalized version of the Ginzburg–Landau equation, and use it to establish where one-dimensional periodic patterns are stable to long-wave perturbations. For most of the calculations we set μp=0, but we do consider the effect of the Proctor term briefly in §3.3.

3.1 Derivation of a generalized Ginzburg–Landau equation

For the model PDE (2.5), we consider only situations where we have a small perturbation of the quartic marginal stability curve. We focus on the case where the perturbed marginal stability curve has two minima, below the cusp (solid lines) in Figure 1. At the quartic minimum, an appropriate scaling is that if the solution u is of O(ϵ), then the slow time T and long length X should be O(ϵ2) and O(ϵ1/2) respectively: the long length X is even longer than the O(ϵ) scaling in the case of a quadratic minimum in (1.2). We therefore restrict f1 and f2 such that the two minima in the marginal stability curve lie in an O(ϵ1/2)×O(ϵ2) box, as illustrated in Figure 3. Other scalings are possible.

Scalings used for the asymptotic analysis. The deviations from a quartic minimum are contained in a $\mathcal{O}(\epsilon ^{1/2}) \times \mathcal{O}(\epsilon ^{2})$ box, and the whole curve is shifted by an $\mathcal{O}(\epsilon )$ amount.
Fig. 3.

Scalings used for the asymptotic analysis. The deviations from a quartic minimum are contained in a O(ϵ1/2)×O(ϵ2) box, and the whole curve is shifted by an O(ϵ) amount.

To simplify the form of the marginal stability curve, we take μp=0 and consider the two minima and intermediate maximum of the function μ=p(k2) in (2.1). The condition for an extremum is p(K)=0, with K=k2: writing 14p(K) as (k2k12)(k2k22)(k2k32), with k1<k2<k3, leads to the conclusion that
(3.1)
To satisfy these three equations, we introduce new parameters γ and δ, defined by
(3.2)
such that the first equation in (3.1) is satisfied. Requiring k1<k2<k3 implies δ>6|γ|0. Substituting (3.2) into the second and third equations in (3.1) yields
These are rearranged to result in expressions for f1, f2 and f1+2f2 in terms of γ and δ:
The restriction δ>6|γ| means that δ24γ2>0, and so f1+2f2=0 implies γ=0 and k2=1.
Now, if the minima and maxima of the marginal stability curve are to satisfy the scalings in Figure 3, we need δ=O(ϵ1/2) and γ no larger than this. We also need p(k22)p(k12) and p(k22)p(k32) to be O(ϵ2):
so with δ=O(ϵ1/2) and γ no larger, these differences fit within O(ϵ2). The overall marginal stability curve is shifted up by an amount p(1)12(12γ2+δ2)=O(ϵ), again within the scalings indicated in Figure 3. The scaling of γ and δ imply f1 and f2 are both O(ϵ), but that f1+2f2=O(ϵ3/2).
We are now in a position to perform the multiple scales analysis. The full set of scalings used are
(3.3)
where ϵμ1 is the amount by which the marginal stability curve is shifted. We also define a singular linear operator L to be (1+x2)4, with Le±ix=0.
Inserting these scalings into (2.5), we obtain at leading order O(ϵ):
which is satisfied by taking
(3.4)
where c.c. represents the complex conjugate. It is useful to observe that x2u1=u1.
Proceeding to O(ϵ3/2), we have
The most convenient way to solve this is to set u3/2=0.
At O(ϵ2), we have
The terms proportional to eix have a prefactor of μ1(f1+f2) We need to eliminate these terms in order to solve for u2, which we do by setting μ1=f1+f2. With this, the remaining terms are 0=Lu2+n2u12, which can be solved to give
The factors 181 come from Le2ix=81e2ix when L is inverted.
At O(ϵ5/2) we have
The only terms involving e±ix on the RHS are the two on the end that combine to have a prefactor of f1+2f2. As discussed above, this combination is a factor of ϵ1/2 smaller than f1 and f2 separately, so this term can be pushed to O(ϵ3) and dropped from this equation. With this, the two remaining terms in the equation can be solved for u5/2:
There is no constant term in u5/2.
Finally, continuing to O(ϵ3), and including the term pushed from O(ϵ5/2), we have
The solvability condition requires the elimination of all terms proportional to e±ix. There are no contributions from Lu3 and from the terms linear in u2 and u5/2. Recalling that x2u1=u1, the terms proportional to eix result in the solvability condition:
(3.5)
where ν1=f1+2f2 and ν2=f1+6f2. The nonlinear term is nA|A|2A, with nA defined in (2.3). This is a generalization of the Ginzburg–Landau equation (1.3), and each term in the equation is, in terms of the original unscaled variables, of order O(ϵ3). The equation has the symmetry (A,X)(A¯,X), which follows from the xx symmetry of the original problem and (3.4). Similar amplitude equations have been proposed before (Raitt & Riecke, 1995; Riecke, 1990; Riley & Davis, 1989), although not formally derived via an asymptotic expansion, to model other problems with very flat marginal stability curves. As we discuss in §6, setting ν1=0 and ν2<0 results in the complex Swift–Hohenberg equation with real coefficients (Gelens & Knobloch, 2010).

3.2 Nonlinear stability of rolls

We can find roll solutions to (3.5) easily, and use the equation to examine their stability to long wavelength Eckhaus instabilities. We restrict ourselves to the supercritical case, where the coefficient nA of the nonlinear term is negative. We consider a roll solution at slightly off-critical wavenumber, i.e.,
(3.6)
which corresponds to a solution u=Rexp(i(1+ϵ1/2q)x) of (2.5) since X=ϵ1/2x, and q=O(1). These are also known as phase-winding solutions. Substituting (3.6) into (3.5), and rearranging, we obtain
(3.7)
where nA<0 for a supercritical bifurcation. The existence boundary of rolls, which is also the marginal stability curve, is where R2=0, or equivalently
(3.8)
with ν1 and ν2 here playing the roles of unfolding parameters for the quartic minimum of the marginal stability curve.
To determine the stability of these roll solutions, we perturb (3.6), writing
where |r|,|ϕ|1, following Hoyle (2006). We substitute this expressions into (3.5), linearize and separate the real and imaginary parts to obtain two linear constant coefficient PDEs for r and ϕ. To solve these, we seek solutions of the form eσT+imX, with mq and derive a quadratic equation for the growth rate σ. One root of this is always negative for a supercritical solution, and the other root is
See Bentley (2012) for details.
For the rolls to be unstable, we require σ>0 for some m, giving a stability boundary defined by
and, recalling (3.7), we find the Eckhaus stability boundary
(3.9)
Figure 4 shows an example of the marginal (3.8) and Eckhaus (3.9) curves. Inside the Eckhaus (dashed) curve, patterns are stable to long wavelength disturbances. Note that ν2<0 when there are two distinct minima, so (3.9) has two vertical asymptotes at q=±ν2/96. Between these asymptotes, μ2 is below the marginal stability curve.
Marginal stability curve (solid) and Eckhaus curves (dashed) for $\nu _{1} = 0$ and $ \nu _{2} = -0.28$. The regions above the dashed curves are Eckhaus stable.
Fig. 4.

Marginal stability curve (solid) and Eckhaus curves (dashed) for ν1=0 and ν2=0.28. The regions above the dashed curves are Eckhaus stable.

3.3 The Proctor term

The presence of these asymptotes means that the Eckhaus boundary does not close up in the middle, as might be expected. However, Proctor (1991) considered an amplitude equation similar to (3.5) but with the first term on the RHS replaced by μ2(1±iX)A, and found that (depending on parameters) the inner edges of the left and right stability boundaries can meet, or the stability region can close with increasing μ with two separate stability regions, and the stability boundaries can be non-monotonic – see Proctor (1991) and Bentley (2012) for examples.

However, when we considered the model equation (2.4), with the Proctor term included, we encountered difficulties with the weakly nonlinear analysis. Writing μ=ϵμ1+ϵ2μ2, as in (3.3), turns out to be inadequate, so we modified the scaling to include an additional ϵ3/2μ3/2(1+x2) term, aiming to relate μ3/2 to μp. It turns out that this does not work either, and possibly further (potentially higher order) terms would need to be considered for a consistent scaling (Bentley, 2012).

4. Lyapunov functional and the first integral

In the Swift–Hohenberg equation, the Lyapunov functional and first integral are useful for finding localized solutions: if two solutions are to be connected by a stationary front, they should have the same values of the first integral and similar values of the Lyapunov functional. In this section, we generalize the SH results to the model equation (2.5).

Multiplying the steady version of (2.5) by ux and integrating by parts gives us a first integral:
Any steady solution of (2.5) must have dHdx=0. It is a straight-forward modification to include the Proctor term from (2.4).
Using the Lyapunov functional for the Swift–Hohenberg equation (1.4) as a starting point, we define a similar functional for (2.5), namely
One can readily show that
It is possible to show that F[u] is bounded below provided f2>4 and n3<0 (Bentley, 2012), and so, as in the SH equation, that stable solutions are local minima of F[u]. Similarly, it is a straight-forward modification to include the Proctor term from (2.4).

Small-amplitude solutions of (2.5) can be found using the generalized GL equation (3.5), and these can be used to find weakly nonlinear estimates of H and F. Alternatively, fully nonlinear solutions of (2.5) can be found numerically, for example by using AUTO (Doedel, 2007). Examples of H and F computed numerically in this way are shown in Figure 5, in the cases where the minima in the marginal stability curve are at the same height, and where there is a single minimum with an almost-minimum just outside the cusp. These were computed using an initial solution on a domain of size L=2π, which was then continued in L, increasing and decreasing to cover the range of wavenumbers for which a pattern solution exists. We note that H and F are both zero at the extremities of the existence region, and that the extrema of H correspond to the Eckhaus stability boundaries.

Plot of $(a)$$30 \mathcal{H}$ and $(b)$$3 \times 10^{3} \mathcal{F}$, against $k$, for $f_{1}=0.14$, $f_{2}=-0.07$ and $\mu =0.07$, computed using AUTO. $(c,d)$: $15 \mathcal{H}$ and $0.5\times 10^{3}\mathcal{F}$, for $f_{1}=0.0488$, $f_{2}=-0.0227$ and $\mu =0.0306$. The red crosses correspond to Eckhaus unstable wavenumbers, and the blue circles to Eckhaus stable wavenumbers. The marginal stability curve (solid black) and Eckhaus curves (black crosses) are also shown – the change from red crosses to blue circles does not exactly match the Eckhaus boundary owing to the scaling of $\mathcal{H}$ and $\mathcal{F}$. Note that in $(c,d)$ there is only one minimum in the marginal stability curve, but there is a region of Eckhaus-stable patterns above the almost-minimum.
Fig. 5.

Plot of (a)30H and (b)3×103F, against k, for f1=0.14, f2=0.07 and μ=0.07, computed using AUTO. (c,d): 15H and 0.5×103F, for f1=0.0488, f2=0.0227 and μ=0.0306. The red crosses correspond to Eckhaus unstable wavenumbers, and the blue circles to Eckhaus stable wavenumbers. The marginal stability curve (solid black) and Eckhaus curves (black crosses) are also shown – the change from red crosses to blue circles does not exactly match the Eckhaus boundary owing to the scaling of H and F. Note that in (c,d) there is only one minimum in the marginal stability curve, but there is a region of Eckhaus-stable patterns above the almost-minimum.

The third tool, mentioned in §1, is the normal form of this variant of the Hamiltonian–Hopf bifurcation with four-fold degenerate eigenvalues ±i. We derive the normal form for this bifurcation in Appendix  A (see equations (A.9) and (A.10)), but as we have not found any first integrals of the normal form, we do not see a way to use it at this point.

5. Localized solutions

The first integral H and the Lyapunov functional F are two imporant tools for identifying where a pattern of one type can be localized within a background of a pattern of another type: the values of the first integral for the two patterns must be the same (since dHdx=0 on any steady solution) and the Lyapunov functional for the two patterns should be approximately the same.

We see from the example in Figure 5(a,b) that there is a range of possible wavenumbers that satisfy the requisite criteria for patterns with two different wavenumbers coexisting, namely there are wavenumbers that have the same value of H, and there are (different) wavenumbers with the same value of F. To narrow down the allowable wavenumbers, we look for wavenumber pairs (k<1,k+>1) such that H(k)=H(k+) and F(k)=F(k+). We do this by looking for intersections of contour lines plotted in the (k+,k) plane. For the parameter values in Figure 5(a,b), a pair of wavenumbers that satisfy this condition are (k+,k)=(1.0778,0.8839). We view such a point as an extension of the Maxwell point for the Swift–Hohenberg equation, though it plays a different role: in the Swift–Hohenberg the localized solutions are organized about the Maxwell point, whereas we use the extension merely as a starting point to look for localized solutions.

On a periodic domain of length L, wavenumbers are restricted to integer multiples of k=2π/L. We therefore construct an initial condition consisting of a region of pattern with wavenumber close to 1.0778 embedded in a background of pattern with wavenumber close to 0.8839. Fixing a domain size L=64×2π, we choose k=58/64 and k+=70/64, and solve (2.5) using a second-order numerical scheme based on Exponential Time Differencing (Cox & Matthews, 2002). One example solution after transients can be seen in Figure 6, which demonstrates that localized solutions to (2.5) exist and are stable. This solution is made up of a high-wavenumber patch (k1.0895) in the centre of the domain, surrounded by low-wavenumber regions (k0.9015). The approximation to the local wavenumber in Figure 6(b) is found via
(5.1)
Numerical simulation of mixed pattern initial condition with wavenumbers $k=58/64=0.90625$ and $k=70/64=1.09375$. The parameter values are: $f_{1}=0.14$, $f_{2}=-0.07$ and $\mu = 0.07$, $n_{2}=0.1$, $n_{3}=-1$, $L=64\times 2\pi $ and timestep $0.01$. $(a)$ final solution profile $u(x)$. $(b)$ approximation to the local wavenumber, defined in (5.1). The values indicated are $k_{-}=0.9015$ and $k_{+}=1.0895$.
Fig. 6.

Numerical simulation of mixed pattern initial condition with wavenumbers k=58/64=0.90625 and k=70/64=1.09375. The parameter values are: f1=0.14, f2=0.07 and μ=0.07, n2=0.1, n3=1, L=64×2π and timestep 0.01. (a) final solution profile u(x). (b) approximation to the local wavenumber, defined in (5.1). The values indicated are k=0.9015 and k+=1.0895.

and matches (at least approximately) the expected values. The oscillations seen in the amplitude and in the local wavenumber represent a beating between the two constituent wavenumbers k and k+ as the pattern adjusts from one wavenumber to the other.

We can continue this and other solutions we have found in AUTO, continuing in μ to obtain the bifurcation diagram shown in Figure 7. Localized solutions lie on distinct branches that do not join up, created in saddle-node bifurcations. The solution branches corresponding to a pattern of single wavelength k or k+ are included for reference. On localized solutions branches closer to the k branch, more of the domain is filled by the k pattern than the k+ pattern, and vice versa. The localized solution branches extend to values of μ below the value at the local maximum of the marginal stability curve (μ(k2)=0.0699).

Bifurcation diagram for parameter values $f_{1}=0.14$, $f_{2}=-0.07$. The critical value of the bifurcation parameter $\mu $ is $\mu _{c}=0.0687$, and the local maximum of the marginal stability curve occurs at $(k_{2}, \mu (k_{2})) = (1,0.0699)$. The thick black branches correspond to periodic patterns with wavenumbers $k_{-}$, $k_{+}$, and are given for reference. The blue branches correspond to localized solutions. Inset is a magnification of the saddle-nodes.
Fig. 7.

Bifurcation diagram for parameter values f1=0.14, f2=0.07. The critical value of the bifurcation parameter μ is μc=0.0687, and the local maximum of the marginal stability curve occurs at (k2,μ(k2))=(1,0.0699). The thick black branches correspond to periodic patterns with wavenumbers k, k+, and are given for reference. The blue branches correspond to localized solutions. Inset is a magnification of the saddle-nodes.

We find similar disconnected branches of localized solutions even when the local minima in the marginal stability curve are not at the same height, and even just outside the cusp, where there is a single minimum and a second almost-minimum, as in Figure 5(c,d). Solutions in this region rely on there being a region of Eckhaus-stable patterns still present above the almost-minimum, disconnected from the marginal stability curve.

5.1 Interpretation of localized solutions via the amplitude equation

The amplitude equation (3.5) has phase-winding solutions A=ReiqX, with R and q related by (3.7). Previous work on this amplitude equation with ν1=0 (Gelens & Knobloch, 2009, 2010; Raitt & Riecke, 1995, 1997) – the complex Swift–Hohenberg equation with real coefficients – has identified solutions that are combinations of two phase-winding solutions with positive and negative values of q: these are precisely the localized patterns we found in the model PDE (2.5) and shown in Figure 6 (with ν1=f1+2f2=0). Here we extend this interpretation to the case ν10.

In order to find localized solutions, we could develop a first integral and a Lyapunov function for (3.5) and look for pairs of solutions with the same values of the quantities. We reserve this for future work, and instead locate localized solutions of (3.5) by starting with a mixture of two phase-winding solutions with constituent wavenumbers q and q+, and timestepping the PDE. Two examples are shown in Figures 8 (with ν10 and a small value of ν2) and 9 (with ν1=0 and a larger value of ν2).

Solution to (3.5), with $\nu _{1}=0.0034$, $\nu _{2}=-0.0874$, $\mu _{2}=0.004$, $n_{2}=0.1$ and $n_{3}=-1$. $(a)$: the real (solid) and imaginary (dashed) parts of the amplitude $A$, $(b)$: the absolute value $|A|$, $(c)$: the reconstructed solution $u = A e^{i x} + \bar{A} e^{-i x}$, and $(d)$: an approximation to the local wavenumber of the reconstructed solution in $(c)$. The dashed lines correspond to $(k_{-},k_{+}) = (0.946,1.053)$.
Fig. 8.

Solution to (3.5), with ν1=0.0034, ν2=0.0874, μ2=0.004, n2=0.1 and n3=1. (a): the real (solid) and imaginary (dashed) parts of the amplitude A, (b): the absolute value |A|, (c): the reconstructed solution u=Aeix+A¯eix, and (d): an approximation to the local wavenumber of the reconstructed solution in (c). The dashed lines correspond to (k,k+)=(0.946,1.053).

Solution to (3.5), with $\nu _{1}=0$, $\nu _{2}=-0.28$, $\mu _{2}=0.0025$, $n_{2}=0.1$ and $n_{3}=-1$. $(a)$: the real (solid) and imaginary (dashed) parts of the amplitude $A$, $(b)$: the absolute value $|A|$, and $(c)$: the reconstructed solution $u = A e^{i x} + \bar{A} e^{-i x}$, with the amplitude $2 |A|$ plotted also.
Fig. 9.

Solution to (3.5), with ν1=0, ν2=0.28, μ2=0.0025, n2=0.1 and n3=1. (a): the real (solid) and imaginary (dashed) parts of the amplitude A, (b): the absolute value |A|, and (c): the reconstructed solution u=Aeix+A¯eix, with the amplitude 2|A| plotted also.

Figure 8(a) shows the real (solid) and imaginary (dashed) parts of the solution A, (A) and (A) respectively. We can clearly see the transition from q to q+ from (A), as the peaks of (A) shift from being on the right of the peaks of (A) to the left, and then back again. Figure 8(b) shows |A|, (half) the amplitude of the reconstructed pattern u=Aeix+A¯eix shown in Figure 8(c). Figure 8(d) shows the approximation to the local wavenumber of the reconstructed pattern.

The example in Figure 9, with ν1=0, is similar, but shows more pronounced beating between the two wavenumbers, evident in the reconstructed solution and in Figure 6. The localized patterns found in the model PDE, with two wavenumbers, are thus interpreted in terms of combinations of phase-winding solutions of the complex Swift–Hohenberg equation, with positive and negative q. Kozyreff et al. (2009) instead interpreted localized patches of patterns with two wavenumbers in terms of localized solutions of the real Swift–Hohenberg equation (1.1). We discuss this in more detail in §6.

5.2 Addition of the Proctor term

The branches of localized solutions shown in Figure 7 (without the Proctor term) do not close. We now investigate the addition of the Proctor term into the model, considering (2.4) with a rather large value of μp=0.65, chosen as to make the effects of this term more pronounced. We also fix f1=0.2814 and f2=0.0721 so that the marginal stability curve has two minima at different heights.

Seeking localized solutions, we plotted (as before) H and F (modified to include the Proctor term), looking for zero contours of H(k+)H(k) and F(k+)F(k), but we found no intersection of contours for the two minima at different heights for this choice of parameters. Notwithstanding this, we returned to wavenumbers k=58/64 and k+=70/64 and used a localized solution constructed from these constituent wavenumbers as a starting point for timestepping and continuation. Part of the resultant localized solution branch is shown in Figure 10. We see that including the Proctor term allows localized solutions of different widths on the same branch, rather than lying on distinct branches as in Figure 7. The existence of the localized solutions is also limited to a finite range of μ values; we expect the upper limit is introduced owing to the band of Eckhaus stable wavenumbers closing.

Branch of localized solutions with $\mu _{p}=-0.65$, $f_{1}=0.2814$, $f_{2}=-0.0721$. The critical value of the bifurcation parameter $\mu $ is $\mu _{c}=-0.314$. The interior saddle-nodes are magnified and the labels $(a)-(f)$ correspond to the solution profiles shown in Figure 11.
Fig. 10.

Branch of localized solutions with μp=0.65, f1=0.2814, f2=0.0721. The critical value of the bifurcation parameter μ is μc=0.314. The interior saddle-nodes are magnified and the labels (a)(f) correspond to the solution profiles shown in Figure 11.

Figure 11 shows solutions at the saddle-nodes indicated in Figure 10. We notice that at each of these saddle-nodes the proportion of each pattern in the domain varies. Lower down the branch the pattern with the smaller wavenumber fills more of the domain, and conversely higher up the branch the pattern with the larger wavenumber fills more of the domain. This behaviour is qualitatively similar to the snaking behaviour of localized solutions in the subcritical Swift–Hohenberg equation (Burke & Knobloch, 2006), in that moving up the snaking branch adds to the width of the spatially periodic part of the localized solution—but the details, with saddle-node bifurcations appearing at many different places along the branch, are considerably more complicated, typical of snaking in more complicated situations such as hexagons (Lloyd et al., 2008) or quasipatterns (Subramanian et al., 2018) in two dimensions.

Solutions at the saddle-nodes indicated in Figure 10.
Fig. 11.

Solutions at the saddle-nodes indicated in Figure 10.

6. The Lugiato–Lefever and complex Swift–Hohenberg equations

As a related problem, we consider the Lugiato–Lefever equation
(6.1)
which governs the envelope of the complex electromagnetic field ψ(t,τ) inside a photonic crystal fibre cavity (Tlidi et al., 2007). In this equation, S represents an injected field, is a cavity detuning, and B2 and B4 incorporate chromatic dispersion. The two time variables t and τ represent, respectively, the average evolution of ψ over one cavity round trip and the fast variations of ψ. We note the similarities between (6.1) and (2.5): the independent variable τ is equivalent to x, there is a cubic nonlinearity, the inhomogeneous S term breaks the ψψ symmetry and implies a quadratic nonlinearity, and a fourth order complex equation is equivalent to an eighth order real equation.

This system also allows marginal stability curves with a double minimum, and Kozyreff et al. (2009) derived an amplitude equation similar to (3.5) for this case. Their analysis concentrates only on the case when the two minima occur at the same height, equivalent to γ=0 in (3.2). This is a degenerate situation however; to recover the generic situation it is necessary to include a B33ψ/τ3 term in (6.1). The degenerate case is chosen by Kozyreff et al. (2009) both as a means of simplifying the analysis and as a situation easily achievable experimentally.

With γ=0, Bentley (2012) showed that the appropriate amplitude equation for the Lugiato–Lefever equation is (3.5) but with ν1=0. This special case is interesting: with ν1=0 and for ν2<0, equation (3.5) is equivalent to the complex Swift–Hohenberg equation:
(6.2)
having scaled and changed notation. This differs from the real SH equation (1.1) in that there is no quadratic term and that the cubic nonlinearity is |A|2A rather than u3. It also differs from the more usual complex SH equation, which has complex coefficients (Sakaguchi, 1997). The complex SH equation with real coefficients has been investigated by Gelens & Knobloch (2010) and is equivalent to to the equation studied by Raitt & Riecke (1995). However, Kozyreff et al. (2009) found the real SH equation (1.1) as the amplitude equation for the Lugiato–Lefever problem, rather than the complex Swift–Hohenberg equation with real coefficients (6.2). This difference arises because effectively Kozyreff et al. (2009) took u1=A(X,T)cos(x) in (3.4) as the solution to the leading order linear equation, with real A. If instead they had taken u1=A(X,T)eix+c.c., with complex A, they would have recovered (6.2). As a result, we believe that the interpretation by Kozyreff et al. (2009) of localized solutions of (6.1) in terms of localized solutions of (1.1) should rather be done in terms of localized solutions of (6.2), which are different and have different stability properties. One feature not captured by localized solutions of (1.1) is the change in wavenumber between the different regions, as discussed in §5. The complex Swift–Hohenberg equation (6.2) does (since the coefficients are real) admit real solutions proportional to cosx, but these solutions are unstable for the parameter values we tried. In addition, once the two minima have different heights, which would happen if the B33ψ/τ3 term were included in the Lugiato–Lefever equation (6.1), then u1=A(X,T)cosx is not a viable starting point: equation (3.5) does not have real solutions when ν10.

7. Discussion and conclusions

The aim of this paper was to develop a new model equation (2.4) that captures qualitatively the behaviour of pattern-forming problems with a quartic marginal stability curve, and then explore the existence (and snaking) of localized solutions within this model equation. While we have focused on a particular example, we expect that the analysis could be extended to other models with a quartic marginal stability curve. There has been much progress in recent years developing a framework for the understanding of localized solutions in the Swift–Hohenberg equation (Dawes, 2010; Knobloch, 2015), where the existence of localized patterns is interpreted in terms of a stable pattern existing at the same parameter values as the stable trivial solution. Here we have shown that this scenario holds in the case of the unfolding of a quartic minimum, where there are coexisting stable patterns with two similar wavenumbers, and localized solutions consisting of combinations of these. This work fits in with other recent efforts that use Swift–Hohenberg-based (and other) models to explore problems with localization and snaking with multiple coexisting patterns (Alrihieli et al., 2021; Knobloch et al., 2019; Subramanian et al., 2021).

We computed the weakly nonlinear amplitude equation for the model (2.5) and recovered a generalized Ginzburg–Landau equation (3.5). This allowed us to compute small-amplitude solutions with different wavelengths and to classify the Eckhaus stable patterns. Our work generalizes that of Raitt & Riecke (1995) and Gelens & Knobloch (2010), and corrects that of Kozyreff et al. (2009), to the case where the heights of the minima in the marginal stability curve are different. We made use of a first integral and a Lyapunov function to identify candidate initial conditions for finding localized solutions. Once the Proctor term, which allows the wavenumber of maximum growth rate to depend on the bifurcation parameter, is included in (2.4), the branches of localized solutions join up, but the snaking we find is considerably more complicated that the standard Swift–Hohenberg scenario.

An alternative approach to looking for localized solutions is to use spatial dynamics, seeking only steady solutions of the model equation. For the Swift–Hohenberg equation, this entails performing a normal form analysis of the Hamiltonian–Hopf bifurcation, which describes the bifurcation from the basic state, and contains the familiar homoclinic snaking of localized solutions (Woods & Champneys, 1999). In this framework, the existence of localized solutions is determined by means of a geometric argument, whereby two integrals of the normal form define a space that can be divided into regions that allow or preclude the existence of localized solutions. In Appendix  A, we derive a normal form for the bifurcation occurring at the quartic minimum of the model equation, following the derivation of the normal form for the Hamiltonian–Hopf bifurcation (Iooss & Adelmeyer, 1998; Woods & Champneys, 1999). We use two different methods (following Iooss & Adelmeyer (1998) and Burke & Knobloch (2007a)) to find the coefficients in the normal form. The two methods do not produce the same values for the coefficients in the normal form, and in either case, there were terms appearing in the normal form that one would have expected to be of higher order. As a consequence of these terms we were unable to find normal form integrals, including those one might expect by extension from the Hamiltonian–Hopf analysis. This means we can not employ a similar geometric analysis to find localized solutions. This normal form analysis merits further investigation. In particular, it would be interesting to ascertain the reasons for the difference between the normal form coefficients calculated by the two different methods, and also to determine whether any variant of the normal form is integrable: the variant we have derived does not seem to be.

Having found numerical localized solutions in the model equation, we could ask next whether examples more closely connected to reality might also have them, for example the magnetized Taylor–Couette system (Stefani et al., 2009) and rotating magnetoconvection (Chandrasekhar, 1961; Cox & Matthews, 2001), where there are suggestions that marginal stability curves can change from having one to having two minima. This last example offers the interesting possibility of exploring two-dimensional patterns, potentially with regions of small hexagons embedded in a background of large hexagons, for example. This would be a natural extension to the study of localized solutions in (variants of) the two-dimensional Swift–Hohenberg equation (Lloyd et al., 2008; Subramanian et al., 2018, 2021), in systems with parametric driving like the Faraday wave experiment (Alnahdi et al., 2018; Arbell & Fineberg, 2000), and in reaction–diffusion systems (Tlidi et al., 2012).

The coalescence of the two minima is in itself an interesting problem. In one-dimensional systems with marginal stability curves with two minima far apart, the natural approach is to reduce the problem to two coupled second-order Ginzburg–Landau equations (Dawes & Proctor, 2008), and there are model equations based on the Swift–Hohenberg equation that allow pattern formation on two length scales (Lifshitz & Petrich, 1997; Müller, 1994; Rucklidge et al., 2012). It would thus be interesting to investigate how one would transition from the situation of two well separated length scales to the unfolding of a quartic minimum, in one or more dimensions.

Acknowledgements

Thomas Wagenknecht (1974–2012) co-supervised much of the work presented in this paper: DCB and AMR gratefully acknowledge his perceptive insight and infectious enthusiasm. This project was inspired by conversations with Rainer Hollerbach. We also acknowledge useful discussions and correspondence with Jonathan Dawes, Gérard Iooss, Edgar Knobloch, Gregory Kozyreff, Eric Lombardi, Michael Proctor and Priya Subramanian. This research was supported by a PhD studentship from the Science and Technology Facilities Council (STFC).

Footnotes

This paper is dedicated to the memory of Thomas Wagenknecht (1974–2012).

References

Alnahdi
,
A.
,
Niesen
,
J.
&
Rucklidge
,
A. M.
(
2018
)
Localized patterns in periodically forced systems: II. Patterns with nonzero wavenumber
.
SIAM J. Appl. Dynam. Syst.
,
17
,
1478
1502
.

Alrihieli
,
H.
,
Rucklidge
,
A. M.
&
Subramanian
,
P.
(
2021
)
Spatial localization beyond steady states in the neighbourhood of the Takens–Bogdanov bifurcation
.
IMA J. Appl. Math.
,
in press
.

Arbell
,
H.
&
Fineberg
,
J.
(
2000
)
Temporally harmonic oscillons in Newtonian fluids
.
Phys. Rev. Lett.
,
85
,
756
759
.

Beck
,
M.
,
Knobloch
,
J.
,
Lloyd
,
D. J.
,
Sandstede
,
B.
&
Wagenknecht
,
T.
(
2009
)
Snakes, ladders, and isolas of localized patterns
.
SIAM J. Math. Anal.
,
41
,
936
972
.

Bentley
,
D. C.
(
2012
)
Localised solutions in the magnetorotational Taylor–Couette flow with a quartic marginal stability curve
, Ph.D. Thesis.
University of Leeds
.

Bortolozzo
,
U.
,
Clerc
,
M. G.
,
Falcon
,
C.
,
Residori
,
S.
&
Rojas
,
R.
(
2006
)
Localized states in bistable pattern-forming systems
.
Phys. Rev. Lett.
,
96
, 214501.

Burke
,
J.
&
Dawes
,
J. H. P.
(
2012
)
Localized states in an extended Swift–Hohenberg equation
.
SIAM J. Appl. Dyn. Syst.
,
11
,
261
284
.

Burke
,
J.
&
Knobloch
,
E.
(
2006
)
Localized states in the generalized Swift–Hohenberg equation
.
Phys. Rev. E
,
73
, 056211.

Burke
,
J.
&
Knobloch
,
E.
(
2007a
)
Normal form for spatial dynamics in the Swift–Hohenberg equation
.
Discrete Contin. Dyn. Sys. Suppl.
,
2007
,
170
180
.

Burke
,
J.
&
Knobloch
,
E.
(
2007b
)
Snakes and ladders: localized states in the Swift–Hohenberg equation
.
Phys. Lett. A
,
360
,
681
688
.

Castelino
,
J. K.
,
Ratliff
,
D. J.
,
Rucklidge
,
A. M.
,
Subramanian
,
P.
&
Topaz
,
C. M.
(
2020
)
Spatiotemporal chaos and quasipatterns in coupled reaction–diffusion systems
.
Phys. D
,
407
, 132475.

Chandrasekhar
,
S.
(
1961
)
Hydrodynamic and Hydromagnetic Stability
.
Oxford
:
Clarendon Press
.

Chapman
,
S. J.
&
Kozyreff
,
G.
(
2009
)
Exponential asymptotics of localised patterns and snaking bifurcation diagrams
.
Phys. D
,
238
,
319
354
.

Coullet
,
P.
,
Riera
,
C.
&
Tresser
,
C.
(
2000
)
Stable static localized structures in one dimension
.
Phys. Rev. Lett.
,
84
,
3069
3072
.

Cox
,
S. M.
&
Matthews
,
P. C.
(
2001
)
New instabilities in two-dimensional rotating convection and magnetoconvection
.
Phys. D
,
149
,
210
229
.

Cox
,
S. M.
&
Matthews
,
P. C.
(
2002
)
Exponential time differencing for stiff systems
.
J. Comp. Phys.
,
176
,
430
455
.

Crawford
,
C.
&
Riecke
,
H.
(
1999
)
Oscillon-type structures and their interaction in a Swift–Hohenberg model
.
Phys. D
,
129
,
83
92
.

Cross
,
M. C.
&
Hohenberg
,
P. C.
(
1993
)
Pattern formation outside of equilibrium
.
Rev. Modern Phys.
,
65
,
851
1112
.

Dawes
,
J. H. P.
(
2010
)
The emergence of a coherent structure for coherent structures: localized states in nonlinear systems
.
Phil. Trans. Roy. Soc. A
,
368
,
3519
3534
.

Dawes
,
J. H. P.
&
Proctor
,
M. R. E.
(
2008
)
Secondary Turing-type instabilities due to strong spatial resonance
.
Proc. Roy. Soc. A
,
464
,
923
942
.

Doedel
,
E. J.
(
2007
) AUTO07p: continuation and bifurcation software for ordinary differential equations.
Technical Report
.
Montreal
:
Concordia University
.

Eckhaus
,
W.
(
1965
)
Studies in Nonlinear Stability Theory
.
Berlin
:
Springer-Verlag
.

Edwards
,
W. S.
&
Fauve
,
S.
(
1994
)
Patterns and quasi-patterns in the Faraday experiment
.
J. Fluid Mech.
,
278
,
123
148
.

Elphick
,
C.
,
Tirapegui
,
E.
,
Brachet
,
M. E.
,
Coullet
,
P.
&
Iooss
,
G.
(
1987
)
A simple global characterization for normal forms of singular vector fields
.
Phys. D
,
29
,
95
127
.

Gelens
,
L.
&
Knobloch
,
E.
(
2009
)
Faceting and coarsening dynamics in the complex Swift–Hohenberg equation
.
Phys. Rev. E
,
80
, 0462201.

Gelens
,
L.
&
Knobloch
,
E.
(
2010
)
Coarsening and frozen faceted structures in the complex Swift–Hohenberg equation
.
Eur. Phys. J. D
,
59
,
23
36
.

Haragus
,
M.
&
Iooss
,
G.
(
2011
)
Local Bifurcations, Center Manifolds, and Normal Forms in Infinite Dimensional Systems
.
London
:
Springer
.

Hegseth
,
J.
,
Vince
,
J. M.
,
Dubois
,
M.
&
Berge
,
M.
(
1992
)
Pattern domains in Rayleigh–Bénard slot convection
.
Europhys. Lett.
,
17
,
413
418
.

Hilali
,
M. F.
,
Métens
,
S.
,
Borckmans
,
P.
&
Dewel
,
G.
(
1995
)
Pattern selection in the generalized Swift–Hohenberg model
.
Phys. Rev. E
,
51
,
2046
2052
.

Holl
,
M. P.
,
Archer
,
A. J.
&
Thiele
,
U.
(
2021
)
Efficient calculation of phase coexistence and phase diagrams: application to a binary phase-field crystal model
.
J. Phys.: Condens. Matter
,
33
, 115401.

Hoyle
,
R. B.
(
2006
)
Pattern Formation: An Introduction to Methods
.
Cambridge
:
Cambridge University Press
.

Iooss
,
G.
&
Adelmeyer
,
M.
(
1998
)
Topics in Bifurcation Theory and Applications. Advanced Series in Nonlinear Dynamics, vol. 3
, 2nd edn.
Singapore
:
World Scientific Publishing Co., Inc.

Iooss
,
G.
&
Pérouème
,
M. C.
(
1993
)
Perturbed homoclinic solutions in reversible 1:1 resonance vector fields
.
J. Diff. Eq.
,
102
,
62
88
.

Knobloch
,
E.
(
1990
)
Pattern selection in long-wavelength convection
.
Phys. D
,
41
,
450
479
.

Knobloch
,
E.
(
2015
)
Spatial localization in dissipative systems
.
Annu. Rev. Condens. Matter Phys.
,
6
,
325
359
.

Knobloch
,
E.
,
Uecker
,
H.
&
Wetzel
,
D.
(
2019
)
Defectlike structures and localized patterns in the cubic-quintic-septic Swift–Hohenberg equation
.
Phys. Rev. E
,
100
, 012204.

Kozyreff
,
G.
,
Tlidi
,
M.
,
Mussot
,
A.
,
Louvergneaux
,
E.
,
Taki
,
M.
&
Vladimirov
,
A. G.
(
2009
)
Localized beating between dynamically generated frequencies
.
Phys. Rev. Lett.
,
102
, 043905.

Kozyreff
,
G.
&
Chapman
,
S. J.
(
2006
)
Asymptotics of large bound states of localized structures
.
Phys. Rev. Lett.
,
97
, 044502.

Lifshitz
,
R.
&
Petrich
,
D. M.
(
1997
)
Theoretical model for Faraday waves with multiple-frequency forcing
.
Phys. Rev. Lett.
,
79
,
1261
1264
.

Lloyd
,
D. J. B.
,
Sandstede
,
B.
,
Avitabile
,
D.
&
Champneys
,
A. R.
(
2008
)
Localized hexagon patterns of the planar Swift–Hohenberg equation
.
SIAM J. Appl. Dynam. Syst.
,
7
,
1049
1100
.

Mamatsashvili
,
G.
,
Stefani
,
F.
,
Hollerbach
,
R.
&
Rüdiger
,
G.
(
2019
)
Two types of axisymmetric helical magnetorotational instability in rotating flows with positive shear
.
Phys. Rev. Fluids
,
4
, 103905.

Müller
,
H. W.
(
1994
)
Model equations for two-dimensional quasipatterns
.
Phys. Rev. E
,
49
,
1273
1277
.

Murdock
,
J.
(
2003
)
Normal Forms and Unfoldings for Local Dynamical Systems
.
New York
:
Springer-Verlag
.

Pomeau
,
Y.
(
1986
)
Front motion, metastability and subcritical bifurcations in hydrodynamics
.
Phys. D
,
23
,
3
11
.

Proctor
,
M. R. E.
(
1991
)
Instabilities of roll-like patterns for degenerate marginal curves
.
Phys. Fluids
,
3
,
299
302
.

Raitt
,
D.
&
Riecke
,
H.
(
1995
)
Domain structures in fourth-order phase and Ginzburg–Landau equations
.
Phys. D
,
82
,
79
94
.

Raitt
,
D.
&
Riecke
,
H.
(
1997
)
Parametric forcing of waves with non-monotonic dispersion relation: domain structures in ferrofluids
.
Phys. Rev. E
,
55
,
5448
5454
.

Rees
,
D. A. S.
&
Mojtabi
,
A.
(
2013
)
The effect of conducting boundaries on Lapwood–Prats convection
.
Int. J. Heat Mass Transfer
,
65
,
765
778
.

Riecke
,
H.
(
1990
)
Stable wave-number kinks in parametrically excited standing waves
.
Europhys. Lett. (EPL)
,
11
,
213
218
.

Riley
,
D. S.
&
Davis
,
S. H.
(
1989
)
Eckhaus instabilities in generalized Landau–Ginzburg equations
.
Phys. Fluids A
,
1
,
1745
1747
.

Rucklidge
,
A. M.
,
Silber
,
M.
&
Skeldon
,
A. C.
(
2012
)
Three-wave interactions and spatiotemporal chaos
.
Phys. Rev. Lett.
,
108
, 074504.

Rucklidge
,
A. M.
&
Silber
,
M.
(
2009
)
Design of parametrically forced patterns and quasipatterns
.
SIAM J. Appl. Dynam. Syst.
,
8
,
298
347
.

Sakaguchi
,
H.
(
1997
)
Standing wave patterns for the complex Swift–Hohenberg equation
.
Prog. Theor. Phys.
,
98
,
577
585
.

Skeldon
,
A. C.
&
Rucklidge
,
A. M.
(
2015
)
Can weakly nonlinear theory explain Faraday wave patterns near onset?
 
J. Fluid Mech.
,
777
,
604
632
.

Spina
,
A.
,
Toomre
,
J.
&
Knobloch
,
E.
(
1998
)
Confined states in large-aspect-ratio thermosolutal convection
.
Phys. Rev. E
,
57
,
524
545
.

Stefani
,
F.
,
Gerbeth
,
G.
,
Gundrum
,
T.
,
Hollerbach
,
R.
,
Priede
,
J.
,
Rüdiger
,
G.
&
Szklarski
,
J.
(
2009
)
Helical magnetorotational instability in a Taylor–Couette flow with strongly reduced Ekman pumping
.
Phys. Rev. E
,
80
, 066303.

Subramanian
,
P.
,
Archer
,
A. J.
,
Knobloch
,
E.
&
Rucklidge
,
A. M.
(
2018
)
Spatially localized quasicrystalline structures
.
New J. Phys.
,
20
, 122002.

Subramanian
,
P.
,
Archer
,
A. J.
,
Knobloch
,
E.
&
Rucklidge
,
A. M.
(
2021
)
Snaking without subcriticality: grain boundaries as non-topological defects
.
IMA J. Appl. Math.
,
in press
.

Swift
,
J.
&
Hohenberg
,
P. C.
(
1977
)
Hydrodynamic fluctuations at the convective instability
.
Phys. Rev. A
,
15
,
319
328
.

Tlidi
,
M.
,
Mussot
,
A.
,
Louvergneaux
,
E.
,
Kozyreff
,
G.
,
Vladimirov
,
A. G.
&
Taki
,
M.
(
2007
)
Control and removal of modulational instabilities in low-dispersion photonic crystal fiber cavities
.
Opt. Lett.
,
32
,
662
664
.

Tlidi
,
M.
,
Sonnino
,
G.
&
Bachir
,
M.
(
2012
)
Predicted formation of localized superlattices in spatially distributed reaction-diffusion solutions
.
Phys. Rev. E
,
86
, 045103.

Topaz
,
C. M.
&
Silber
,
M.
(
2002
)
Resonances and superlattice pattern stabilization in two-frequency forced Faraday waves
.
Phys. D
,
172
,
1
29
.

Woods
,
P. D.
&
Champneys
,
A. R.
(
1999
)
Heteroclinic tangles and homoclinic snaking in the unfolding of a degenerate reversible Hamiltonian–Hopf bifurcation
.
Phys. D
,
129
,
147
170
.

A. A normal form for the model equation

In this Appendix we derive the normal form for the model equation, with the intention of extending the analysis of the Hamiltonian–Hopf bifurcation in the Swift–Hohenberg equation (Iooss & Pérouème, 1993; Woods & Champneys, 1999) to this case of quadruply degenerate eigenvalues. Our approach is based on that of these authors (see also Haragus & Iooss (2011)), along with the work of Burke & Knobloch (2007a). For additional details of the calculations in this Appendix, see Bentley (2012).

In general, we consider dynamical systems of the form
(A.1)
where we assume that
and also that there is a symmetry R such that
This symmetry R is known as a reversibility symmetry. We consider reversible systems because of the invariance of the model equation (2.5) under spatial reversibility xx. The parameter vector ϱ=0 is chosen so that the eigenvalues of the z=0 equilibrium are all on the imaginary axis.
We aim to derive the normal form of the model equation near the bifurcation This derivation essentially entails finding a near-identity transform
such that we may write (A.1) in the form
(A.2)
Here Φ(z~,ϱ) and P(z~,ϱ) are (n-dimensional vectors of) polynomials of degree kp, and L0 is a constant coefficient matrix in Jordan normal form. The polynomial P satisfies the so-called homological equation
where L0 is the adjoint (conjugate transpose) of L0. An equivalent statement is found by differentiating the equation with respect to x and evaluating at x=0, which gives
where D~P(z~,ϱ) is the Jacobian matrix of P.

There is some freedom in determining the polynomial P; the idea is to choose P to be as simple as possible. Of course, this freedom means that there is not a unique normal form. Rather, the choice of P is known as the style of the normal form. The two main styles are the inner-product style popularized by Elphick et al. (1987) and the sl(2) style popularized by Cushman and Sanders (Murdock, 2003). We will use the inner-product style, since this style is used for the Hamiltonian–Hopf bifurcation (Iooss & Adelmeyer, 1998), which describes the bifurcation from the basic state in the Swift–Hohenberg equation.

In what follows, we will derive the normal form at the codimension 3 point μ=f1=f2=0, which corresponds to ϱ=0 in (A.1). So, the homological equation we will actually use is
(A.3)
The parameters μ, f1 and f2 can be added in as unfoldings once the normal form has been found.

A.1 Linear part of the normal form

In this section, we write the dynamical system describing steady solutions of the model equation. We then determine the linear part of the coordinate transformation zz~. By considering steady solutions of our model equation (2.5) with μ=f1=f2=0, i.e., solutions of
(A.4)
we can convert (A.4) into a system of eight first-order ODEs, such that we have the appropriate form (A.1). To do this, we introduce new variables
and write
(A.5)
where z=(z1,z2,z3,z4,z5,z6,z7,z8)T. The linear and nonlinear parts of (A.5) are given by
The reversibility R acting on the elements of z is defined as Rzi=(1)i1zi for i=1,,8.
The first step is to transform L0 into Jordan normal form. The eigenvalues are λ±=±i with algebraic multiplicity 4 and geometric multiplicity 1, so each eigenvalue has one eigenvector and three generalized eigenvectors. These are readily found:
(A.6)
with Lζ0=λ+ζ0 and Lζj=λ+ζj+ζj1, with j=1,2,3. We now define the linear transformation
(A.7)
where the overbar denotes complex conjugation, and A, B, C and D are complex functions of x. The transformed linear normal form is dz~dx=L0z~, where the transformed L0 and its adjoint L0 are
Thus, the linear part of the normal form is
(A.8)
along with the complex conjugates of these.

A.2 Nonlinear part of the normal form

To determine the nonlinear part of the normal form P(z~) we make use of the homological equation (A.3). We truncate at cubic order, setting kp=3 in (A.2). We also take in to account the linear transformation in §A.1, now thinking of z~ as (A,B,C,D,A¯,B¯,C¯,D¯).

One possible approach to determining the nonlinear part of the normal form is to suppose P(z~) contains all possible quadratic and cubic combinations of the components of z~, i.e.,
such that i+j+k+l+m+n+o+p=2 or 3. Then, plugging this into the homological equation (A.3) gives the terms and the combinations in which they must appear. This allows us to write the normal form as:
(A.9a)
 
(A.9b)
 
(A.9c)
 
(A.9d)
where the coefficients γi, i=1,,26 are to be determined by transforming the nonlinear term n2u2+n3u3. One may notice that (A.9) contains no quadratic terms; the requirement that we satisfy the homological equation (A.3) excludes them. Had we chosen a different normal form style, e.g., the sl(2) style (Murdock, 2003), then it is possible that (A.9) would have contained quadratic terms.

An alternative approach to constructing the normal form is to find first integrals of the homological equation and construct the polynomials P(z~) using these, following Elphick et al. (1987). For details, see Bentley (2012). The end result is a set of first integrals c1, …, c7, given in Table A.1. Note that c2, c4 and c6 contain log(A), so we also use w1 and w2, which are combinations of the first seven with the log(A) dependence eliminated. We note that the integrals c1, c2 and c3 are also integrals of the characteristic system in the four-dimensional Hamiltonian–Hopf case (Iooss & Adelmeyer, 1998).

Table A1.

The first integrals of the homological equation.

c1AA¯
c2iBA+log(A)
c3i(AB¯A¯B)
c4CAiBAlog(A)12(log(A))2
c5AC¯BB¯+A¯C
c6iDACAlog(A)+iB2A(log(A))2+16(log(A))3
c7i(AD¯BC¯+B¯CA¯D)
w11A2(B22AC)
w2iA2((3ADBC)+BA(B22AC))
c1AA¯
c2iBA+log(A)
c3i(AB¯A¯B)
c4CAiBAlog(A)12(log(A))2
c5AC¯BB¯+A¯C
c6iDACAlog(A)+iB2A(log(A))2+16(log(A))3
c7i(AD¯BC¯+B¯CA¯D)
w11A2(B22AC)
w2iA2((3ADBC)+BA(B22AC))
Table A1.

The first integrals of the homological equation.

c1AA¯
c2iBA+log(A)
c3i(AB¯A¯B)
c4CAiBAlog(A)12(log(A))2
c5AC¯BB¯+A¯C
c6iDACAlog(A)+iB2A(log(A))2+16(log(A))3
c7i(AD¯BC¯+B¯CA¯D)
w11A2(B22AC)
w2iA2((3ADBC)+BA(B22AC))
c1AA¯
c2iBA+log(A)
c3i(AB¯A¯B)
c4CAiBAlog(A)12(log(A))2
c5AC¯BB¯+A¯C
c6iDACAlog(A)+iB2A(log(A))2+16(log(A))3
c7i(AD¯BC¯+B¯CA¯D)
w11A2(B22AC)
w2iA2((3ADBC)+BA(B22AC))
From the first integrals of the homological equation, we can construct the nonlinear part of the normal form. We make the change of variables
and solve the homological equation in these new variables (see Bentley, 2012), giving, for the first equation in the normal form:
some some arbitrary function φ, provided that P1=Aφ is a polynomial in its eight arguments. Here we use w1 and w2 in preference to c4 and c6.

In the derivation of the normal form for the Hamiltonian–Hopf bifurcation, the equivalent equation at this stage is P1(A,B,A¯,B¯)=Aφ(c1,c2,c3). The argument is then that φ is a polynomial in c1 and c3, and independent of c2. This is because of the log dependence of c2: as A0, the logarithmic behaviour of c2 does not match the polynomial behaviour of P1, and thus φ must be independent of c2. This argument follows through to our case as far as c2 is concerned, but there are additional considerations regarding w1 and w2, which have A2 in their denominators. This dependence in Aφ is eliminated by taking certain combinations of w1 and w2: for example, Ac1w1=A¯(B22AC), which is fine, as is A(c1w2+w1c3)=A¯(3ADBC)+B¯(B22AC), while Ac1w2 has an A in the denominator and so is not a polynomial. In fact, only the two combinations c1w1 and c1w2+c3w1 are needed for P1 for dAdx, but additional combinations appear in the other three equations.

After computing these and (re)labelling the arbitrary functions as P, Q, R and S, to be consistent with the notation of the normal form of the Hamiltonian–Hopf bifurcation (Burke & Knobloch, 2007a; Iooss & Adelmeyer, 1998), we have the eight-dimensional normal form
(A.10a)
 
(A.10b)
 
(A.10c)
 
(A.10d)
The functions P, Q, R and S are understood to include only those combinations of w1 and w2 that result in polynomial contributions to the normal form. The nonlinear terms up to cubic order are:
 
 
 
where the Pi, Qi, Ri, Si and Tj, i=1,,4, j=1,,10 are real coefficients (with the prefactor i included in the cases of T2, T4, T5, T6 and T10). These terms may seem somewhat arbitrary, but are in fact very specific combinations to match the terms found by solving the homological equation (compare with (A.9)). The relation between the normal form coefficients here and the normal form coefficients in (A.9) is given in Table A.2.
Table A2.

Relation between the normal coefficients in (A.9) and the normal form coefficients in (A.10).

γ1=iP1γ2=P2γ3=iP3γ4=P4γ5=T2γ6=T1
γ7=Q1γ8=iQ2γ9=Q3γ10=iQ4γ11=T3γ12=T4
γ13=iR1γ14=R2γ15=iR3γ16=R4γ17=T5γ18=2T6
γ19=T7/2γ20=S1γ22=iS2γ22=S3γ23=iS4γ24=T8
γ25=T9γ26=T10
γ1=iP1γ2=P2γ3=iP3γ4=P4γ5=T2γ6=T1
γ7=Q1γ8=iQ2γ9=Q3γ10=iQ4γ11=T3γ12=T4
γ13=iR1γ14=R2γ15=iR3γ16=R4γ17=T5γ18=2T6
γ19=T7/2γ20=S1γ22=iS2γ22=S3γ23=iS4γ24=T8
γ25=T9γ26=T10
Table A2.

Relation between the normal coefficients in (A.9) and the normal form coefficients in (A.10).

γ1=iP1γ2=P2γ3=iP3γ4=P4γ5=T2γ6=T1
γ7=Q1γ8=iQ2γ9=Q3γ10=iQ4γ11=T3γ12=T4
γ13=iR1γ14=R2γ15=iR3γ16=R4γ17=T5γ18=2T6
γ19=T7/2γ20=S1γ22=iS2γ22=S3γ23=iS4γ24=T8
γ25=T9γ26=T10
γ1=iP1γ2=P2γ3=iP3γ4=P4γ5=T2γ6=T1
γ7=Q1γ8=iQ2γ9=Q3γ10=iQ4γ11=T3γ12=T4
γ13=iR1γ14=R2γ15=iR3γ16=R4γ17=T5γ18=2T6
γ19=T7/2γ20=S1γ22=iS2γ22=S3γ23=iS4γ24=T8
γ25=T9γ26=T10

Having found the terms present in the normal form, we now wish to find the coefficients of these terms. In the following sections, we will describe two methods to do this: by solving a linear system of equations derived from the system of ODEs (A.5), following Iooss & Adelmeyer (1998), and an asymptotic scaling method, following Burke & Knobloch (2007a).

A.3 Determining the normal form coefficients I: Nonlinear coordinate transform

We have so far described the linear transformation (A.7). Following Iooss & Adelmeyer (1998), we now add nonlinear terms to the transformation, in particular a polynomial Φ(z~), such that we have
(A.11)
We fix Φ such that it contains only quadratic and cubic terms. Substituting this into (A.5) and matching like powers of the variables results in a relationship between the parameters n2 and n3 in the model (A.4) and the normal form coefficients in (A.10):
(A.12)
All the normal form coefficients are either purely real or purely imaginary, which is a consequence of the reversibility symmetry.

A.4 Determining the normal form coefficients II: Asymptotic scaling method

This method involves expanding both the steady model equation (A.4) and the normal form equations (A.10) in powers of a small parameter ϵ, following Burke & Knobloch (2007a). The normal form coefficients can then be found by matching the equations at each order of ϵ.

A.4.1 Model equation expansion
We introduce a small parameter ϵ, and define a long length scale X=ϵ1/2x, as we did in the weakly nonlinear analysis in §3. We then expand u(x) in terms of this small parameter, i.e.,
(A.13)
The summation index runs from n=2 because the amplitude of u is O(ϵ) (see §3). We need to go to n=12 in order to determine all the coefficients in the normal form (A.10).
Substituting (A.13) into (A.4), we obtain equations to be solved at each order of ϵ1/2, i.e.,
and similarly for higher orders, up to O(ϵ6). This analysis is equivalent to the one performed in §3, though there is a factor of two difference in the subscript of u between the two calculations.
The leading order solution is given by
and similarly from O(ϵ3/2) we have
In the weakly nonlinear analysis in §3, we set u3/2=0 (equivalently u3=0 here); in this analysis we keep u3 non-zero. The difference is of no consequence, however.
Proceeding to O(ϵ2), the ansatz u4=λ4+A4(X)eix+B4(X)e2ix+c.c., where λ4 is real, leads to the solution
and A4, A¯4 are as yet unknown. We have dropped the explicit X dependence for convenience. Again, this is exactly the solution found in §3; similarly the solution at O(ϵ5/2) is as described in §3.
At O(ϵ3), this analysis begins to differ from the one in §3. In particular, whereas in §3 the time derivatives first appeared at O(ϵ3), here we have no time derivatives. Instead we obtain a fourth-order ODE for A2, namely
where the prime denotes differentiation with respect to X. We note that the coefficient of the nonlinear term is nA from (2.3). We obtain similar equations for the Aj, j=3,,8 at O(ϵ2+j/2). For example, we find
at O(ϵ7/2). Continuing this process up to O(ϵ6), we obtain all the required equations. We may then reconstitute these into one equation by defining
The resulting equation has derivatives up to eighth order. However, derivatives higher than fourth order appear at higher order than the A term, and these can be eliminated by repeatedly differentiating the resulting equation and substituting back in to the equation. The resulting equation is then
(A.14)
where the θi, i=1,,3 coefficients are
To match to the scaled normal form equation (derived in the next section), we require one more transformation. This is
(A.15)
where the ρi, i=1,,6 coefficients are to be determined through the matching procedure. Under this transformation, (A.14) becomes
(A.16)
We have continued up to O(ϵ3), but the number of terms in (A.16) quickly escalates to such an extent that it is not instructive to include them all here; the truncated form of (A.16) is sufficient for the argument presented below.
A.4.2 Normal form scaling
To match the scaling in §A.4.1, we write
where, as before, X=ϵx, and we have factored out an eix dependence. With this, the normal form for Ax, (A.10a), becomes
Cancelling the common factor eix, subtracting iϵA~ from both sides and dropping the tildes, we have
(A.17)
We may cast this in appropriate form for the analysis by dividing through by a factor of ϵ3/2 and rearranging, such that we have
(A.18)
We do not explicitly write out all the terms for the sake of clarity in explaining the procedure.
Similarly, from (A.10) we have
(A.19)
and
(A.20)
We also obtain an equation for DX, which we leave in the form
(A.21)
The next step is to differentiate (A.18) with respect to X, giving
(A.22)
Substituting (A.22) into (A.19), we have
(A.23)
This procedure of differentiation and substitution is repeated twice more. Differentiating (A.23) with respect to X we have
(A.24)
Before substituting (A.24) into (A.20), we first replace the B and B¯ terms using (A.18). This yields
(A.25)
Now substituting (A.24) into (A.25) we have
(A.26)
Differentiating one final time, we have
(A.27)
We now notice that we have two equations for DX, (A.21) and (A.27), and thus equating the two will yield an equation for AXXXX. Before we do this, we first should use (A.22) and (A.23) to replace the B, B¯, C and C¯ terms in (A.21). This gives
(A.28)
Equating (A.27) and (A.28), we find
(A.29)

As in §A.4.1, continuing up to O(ϵ3) yields a very large number of terms, so we truncate at the same order as (A.16). This is sufficient for the coefficient matching in the next section.

A.4.3 Order-by-order matching
We now match the two equations (A.16) and (A.29) at each power of ϵ1/2 to recover the normal form coefficients. The subscript X in (A.29) and superscript prime in (A.16) are now equivalent. At leading order the solution is immediately recovered, and we have
This is the same as calculated via the nonlinear coordinate transform method, as given in (A.12).
Proceeding to O(ϵ1/2) we have to solve the equations
The solution to these two equations is
again, the same as (A.12).
At the next order, O(ϵ), we have the four equations
Solving these four equations simultaneously, we find
Comparing with (A.12), we see that there is a disagreement in three out of the four terms. T8 agrees, whereas the other three agree only in the dependence on the cubic nonlinearity coefficient n3.

The differences in Q1, R2 and S3 are 2n22/7, 3n22/7, 2n22/7 respectively. This trend continues at higher orders, i.e., the normal form coefficients found by the asymptotic scaling method agree with those found by the nonlinear coordinate transform method only in the dependence on n3. We note that any extra terms included in the transformation (A.15) to try and rectify this would depend on θ1, and thus change the n3-dependence of the normal form coefficients.

A possible explanation for this discrepancy is the non-uniqueness of the normal form. In the analysis of the Swift–Hohenberg equation, the equivalent linear transformation to (A.7) is more generally a two-parameter family of transformations (Burke & Knobloch, 2007a). Fixing the value of these parameters at linear order determines the values of the normal form coefficients at higher order. By extension, we similarly will have a four-parameter family of transformations. Whereas in the Hamiltonian–Hopf case setting these parameters to determine the linear coordinate transform automatically determines the normal form coefficients at higher order, that is not necessarily the case here. There may be extra parameters that have implicitly been set differently by the two different methods.

Similarly, it is possible the w1 and w2 terms could have introduced an extra hidden parameter into the system, and the value of this parameter chosen by the two methods is not consistent.

A.5 First integrals of the normal form

The normal form for the Hamiltonian–Hopf bifurcation has two integrals that allow one to determine geometrically the solutions of the Swift–Hohenberg equation (Iooss & Adelmeyer, 1998; Iooss & Pérouème, 1993; Woods & Champneys, 1999). So, for the normal form in our case to be of any use, we need to find integrals of the normal form (A.10).

One of the integrals of the normal form for the Hamiltonian–Hopf bifurcation is also an integral of the characteristic system used to derive the normal form. Why this should be the case is not obvious. Nonetheless, we might hope that one of the integrals (or some combination of the integrals) in Table A.1 is also be an integral for the normal form (A.10).

We first concentrate on the linear terms of the normal form (A.8), and consider the integral c7=i(AD¯BC¯+B¯CA¯D). Differentiating c7 with respect to x, we have
using (A.8). Thus c7 is an integral of the linear normal form (A.8). There are three similar integrals of the linear normal form, to wit
 
and
However, none of these four integrals of the linear normal form are integrals of the nonlinear normal form, nor is any combination of them. This failure to extend to the nonlinear normal form is entirely down to the presence of the w1 and w2 terms in the normal form. In particular, it is the fact that w1 and w2 are not real that the extension fails. A wider search for other possible forms for integrals did not uncover any.

The absence of integrals of the normal form prevents the extension of the geometric analysis of the Hamiltonian–Hopf bifurcation to our degenerate situation. A possible explanation for the lack of integrals of the normal form is the non-uniqueness of the normal form: other choices of which terms to keep in the polynomial P in (A.2) might lead to an integrable normal form. It would be interesting to pursue this further.

This is an Open Access article distributed under the terms of the Creative Commons Attribution License (https://creativecommons.org/licenses/by/4.0/), which permits unrestricted reuse, distribution, and reproduction in any medium, provided the original work is properly cited.