This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

A Bifurcation Lemma for Invariant Subspaces

John M. Neuberger Nándor Sieben  and  James W. Swift [email protected], [email protected], [email protected] Department of Mathematics and Statistics, Northern Arizona University PO Box 5717, Flagstaff, AZ 86011-5717, USA
Abstract.

The Bifurcation from a Simple Eigenvalue (BSE) Theorem is the foundation of steady-state bifurcation theory for one-parameter families of functions. When eigenvalues of multiplicity greater than one are caused by symmetry, the Equivariant Branching Lemma (EBL) can often be applied to predict the branching of solutions. The EBL can be interpreted as the application of the BSE Theorem to a fixed point subspace. There are functions which have invariant linear subspaces that are not caused by symmetry. For example, networks of identical coupled cells often have such invariant subspaces. We present a generalization of the EBL, where the BSE Theorem is applied to nested invariant subspaces. We call this the Bifurcation Lemma for Invariant Subspaces (BLIS). We give several examples of bifurcations and determine if BSE, EBL, or BLIS apply. We extend our previous automated bifurcation analysis algorithms to use the BLIS to simplify and improve the detection of branches created at bifurcations.

Key words and phrases:
coupled networks, synchrony, bifurcation, invariant subspaces
2020 Mathematics Subject Classification:
34A34, 34C23, 35J61, 37C79, 37C81

1. Introduction

We present a Bifurcation Lemma for Invariant Subspaces (BLIS). The BLIS proves the existence of bifurcating solution branches of a nonlinear equation F(s,x)=0F(s,x)=0 with a single parameter ss. It applies the Bifurcation from a Simple Eigenvalue (BSE) Theorem of Crandall and Rabinowitz [4] to nested invariant subspaces. The BSE Theorem is a fundamental result well known to researchers in partial differential equations (PDE). The equally fundamental Equivariant Branching Lemma (EBL) [3, 10, 25] is a powerful tool for the study of bifurcations in symmetric dynamical systems. While symmetry in systems leads to invariant fixed point subspaces [12], the structure of some systems, especially coupled networks, causes additional invariant subspaces [11, 14, 15, 17, 21]. In [14], we call these anomalous invariant subspaces. The BLIS can be thought of as extending the EBL from fixed point subspaces of symmetric systems to all invariant subspaces. The BLIS is a generalization of the Synchrony Branching Lemma, which describes some bifurcations from the fully synchronous state in coupled oscillators. See [5, 9, 24] and [11, Section 18.3].

The BLIS predicts the branching of solutions in a wide variety of applications, including all branching predicted by the EBL. The BLIS lends itself to implementation as a numerical algorithm, and provides a bridge between bifurcation theory in PDE and dynamical systems. Figure 1 shows the relationship between BLIS, BSE, and EBL, and references our example applications.

Some invariant subspaces are caused by the symmetry of a dynamical system. The topology of a network of coupled cells can introduce additional invariant subspaces. In general, there can be even more invariant subspaces. Our current paper does not discuss the reason for, or computation of, these invariant subspaces. A search for a more general theory of invariant subspaces can be found in [23]. This work includes the Equivaroid Branching Lemma [23, Theorem 4.2.2], which is similar to our BLIS.

In Section 2 we give some definitions and introduce notation, and then state and prove the BLIS and some related propositions. In Section 3 we present numerous examples of applications of the BLIS. We consider one-dimensional systems, coupled cell networks, and PDE. Lastly, in Section 4 we give some details concerning the algorithms and numerical implementation we used to generate the bifurcation diagrams in the coupled network examples. Algorithms for finding invariant subspaces for small networks are developed in [2, 13, 17, 21], with implementations found at [16, 20]. Our code for branch following, which is freely available at [18], uses the BLIS. It improves the Newton method implementations in [6, 14, 15] which use only the fixed point subspaces of the group action.

EBLBLISBSE3.3(c)3.2(c)3.19(c)3.20(c)3.24(c)3.1(c)3.17(b)3.21(c)3.22(c)3.23(a)3.25(a)3.17(a)3.18(c)3.23(b)3.25(b)3.4(c)3.5(c)3.25(c)
Figure 1. The relationship between the Bifurcation Lemma for Invariant Subspaces, Bifurcation from a Simple Eigenvalue, and the Equivariant Bifurcation Lemma. The numbers indicate examples.

2. The Main Results

In this section, we give some background and state the BSE Theorem of Crandall and Rabinowitz [4]. We then use the BSE Theorem to prove the BLIS.

Definition 2.1.

Let F:I×XXF:I\times X\to X, where II is an open interval and XX is a Banach space. A subspace WW of XX is FF-invariant if F(s,W)WF(s,W)\subseteq W for all sIs\in I.

This extends the standard definition that WW is an FF-invariant subspace for F:XXF:X\to X provided F(W)WF(W)\subseteq W.

Example 2.2.

The trivial subspace {0}X\{0\}\subseteq X is FF-invariant if and only if F(s,0)=0F(s,0)=0 for all sIs\in I. Also, the full subspace XX is FF-invariant for any F:I×XXF:I\times X\to X.

We often get FF-invariant subspaces when there is a symmetry in the system. We recall some definitions involving group actions as they relate to invariant subspaces. Assume a compact Lie group Γ\Gamma acts linearly on a Banach space XX by the function (γ,x)γx:Γ×XX(\gamma,x)\mapsto\gamma x:\Gamma\times X\to X. We say that F:I×XXF:I\times X\to X is Γ\Gamma-equivariant if F(s,γx)=γF(s,x)F(s,\gamma x)=\gamma F(s,x) for all (s,x)I×X(s,x)\in I\times X, γΓ\gamma\in\Gamma. The stabilizer of xx, or the isotropy subgroup of Γ\Gamma with respect to xXx\in X, is Γx:={γΓ:γx=x}\Gamma_{x}:=\{\gamma\in\Gamma:\gamma x=x\}. An isotropy subgroup of Γ\Gamma is Σ=Γx\Sigma=\Gamma_{x} for some xx. A subgroup of Γ\Gamma is not necessarily an isotropy subgroup of the group action. For every subgroup Σ\Sigma of Γ\Gamma, the fixed point subspace of Σ\Sigma is

Fix(Σ):={xX:σx=x for all σΣ}.\operatorname{Fix}(\Sigma):=\{x\in X:\sigma x=x\text{ for all }\sigma\in\Sigma\}.

The point stabilizer of a set UXU\subseteq X is

pStab(U):={γΓ:γx=x for all xU}.\text{pStab}(U):=\{\gamma\in\Gamma:\gamma x=x\mbox{ for all }x\in U\}.

The group action on XX extends to an action on subsets of XX. If WW is an FF-invariant subspace, then γW\gamma W is also an FF-invariant subspace for all γΓ\gamma\in\Gamma. The Γ\Gamma-orbit of an FF-invariant subspace WW is {γW:γΓ}\{\gamma W:\gamma\in\Gamma\}.

Proposition 2.3.

Given a subgroup Σ\Sigma of a compact Lie group Γ\Gamma that acts linearly on a Banach space, Fix(Σ)\operatorname{Fix}(\Sigma) is an FF-invariant subspace if FF is Γ\Gamma-equivariant.

Proof.

First of all, Fix(Σ)\operatorname{Fix}(\Sigma) is a subspace of XX because Γ\Gamma acts linearly. Assume xFix(Σ)x\in\operatorname{Fix}(\Sigma). Then F(s,x)=F(s,σx)=σF(s,x)F(s,x)=F(s,\sigma x)=\sigma F(s,x) for all σΣ\sigma\in\Sigma. Thus F(s,x)Fix(Σ)F(s,x)\in\operatorname{Fix}(\Sigma). ∎

Example 2.4.

Let F:CCF:C^{\infty}\to C^{\infty} be defined by F(u)=u′′F(u)=u^{\prime\prime}, where CC^{\infty} is the space of infinitely differentiable real-valued functions on \mathbb{R}. The group 2={1,γ}\mathbb{Z}_{2}=\{1,\gamma\} acts on CC^{\infty} by (γu)(t)=u(t)(\gamma u)(t)=u(-t). Note that FF is 2\mathbb{Z}_{2}-equivariant because F(γu)(t)=u′′(t)=(γF(u))(t)F(\gamma u)(t)=u^{\prime\prime}(-t)=(\gamma F(u))(t). The FF-invariant fixed point subspace Fix(2)\operatorname{Fix}(\mathbb{Z}_{2}) is the subspace of even functions.

Recall that a simple C1C^{1}-curve is the image of an injective C1C^{1}-function p:(a,b)I×Xp:(a,b)\to I\times X with non-vanishing derivative. Note that a simple C1C^{1}-curve does not cross itself, and it has no cusps or corners.

Definition 2.5.

Let F:I×XXF:I\times X\to X be as in Definition 2.1. A solution branch of FF, or simply a branch, is a simple C1C^{1}-curve that is a subset of F1({0})F^{-1}(\{0\}). A WW-branch is a solution branch contained in I×WI\times W, where WW is an FF-invariant subspace of XX.

Example 2.6.

The function F:×F:\mathbb{R}\times\mathbb{R}\to\mathbb{R} defined by F(s,x)=sxF(s,x)=sx has two solution branches, with equations s=0s=0 and x=0x=0. The L-shaped curve {(s,0)s0}{(0,x)x0}\{(s,0)\mid s\geq 0\}\cup\{(0,x)\mid x\geq 0\} is not a solution branch because it has a corner at (0,0)(0,0).

Definition 2.7.

A point pp in a solution branch of FF is a bifurcation point if every neighborhood of pp in I×XI\times X contains a zero of FF that is not in the solution branch.

Typically, a bifurcation point is the intersection of two or more branches. In the most recent example, (0,0)(0,0) is a bifurcation point. Note that a fold point, sometimes called a saddle-node bifurcation, is not a bifurcation point by this definition, which follows [4]. This definition also ignores Hopf bifurcations.

The following proposition is a consequence of the Implicit Function Theorem, and it gives conditions for which a solution (s,x)(s^{*},x^{*}) to F(s,x)=0F(s,x)=0 can be extended to a unique solution branch within an invariant subspace. We call this curve of solutions the mother branch because we anticipate that the hypotheses of Theorem 2.12 hold.

Proposition 2.8.

Let II be an open interval, XX be a Banach space, and F:I×XXF:I\times X\to X be a C2C^{2}-function. Suppose WmW_{m} is an FF-invariant subspace of XX, (s,x)I×Wm(s^{*},x^{*})\in I\times W_{m} satisfies F(s,x)=0F(s^{*},x^{*})=0, and D2F(s,x)|WmD_{2}F(s^{*},x^{*})|_{W_{m}} is nonsingular. Then there exists a neighborhood UU of (s,x)(s^{*},x^{*}) in I×XI\times X, and a C2C^{2}-function bm:ImWmb_{m}:I_{m}\to W_{m} defined on an open interval containing ss^{*} such that bm(s)=xb_{m}(s^{*})=x^{*} and

F1({0})(I×Wm)U=Cm,F^{-1}(\{0\})\cap(I\times W_{m})\cap U=C_{m},

where

(1) Cm:={(s,bm(s))sIm}.C_{m}:=\{(s,b_{m}(s))\mid s\in I_{m}\}.

We call bmb_{m} the mother branch function and its graph CmC_{m} the mother branch.

Proof.

This is the Implicit Function Theorem applied to the restriction Fm:I×WmWmF_{m}:I\times W_{m}\to W_{m} of FF to I×WmI\times W_{m}. The only subtlety is that (I×Wm)U(I\times W_{m})\cap U is a neighborhood of (s,x)(s^{*},x^{*}) in I×WmI\times W_{m}. ∎

Remark 2.9.

If Wm={0}W_{m}=\{0\} is an invariant subspace, then D2F(s,x)|WmD_{2}F(s^{*},x^{*})|_{W_{m}} is vacuously nonsingular. Here the mother branch function is bm(s)=0b_{m}(s)=0 for all sIs\in I, and the mother branch is called the trivial branch.

The BLIS describes a bifurcation of the mother branch of Proposition 2.8 at (s,x)(s^{*},x^{*}) to a daughter branch. This is a bifurcation from a simple eigenvalue for the restricted function Fd:I×WdWdF_{d}:I\times W_{d}\to W_{d}, where WmWdW_{m}\subsetneq W_{d} are the invariant subspaces corresponding to the mother and daughter branches, respectively. To this end, we first give a slight modification of the BSE Theorem [4, Theorem 1.7]. Motivated by Definition 2.1, we assume F:Im×XXF:I_{m}\times X\to X, while [4] assumes F:(1,1)×XYF:(-1,1)\times X\to Y. In our theorem, the bifurcation is at (s,0)(s^{*},0), whereas [4] assumes the bifurcation point is (0,0)(0,0).

Theorem 2.10 (BSE, Theorem 1.7 of [4]).

Let XX be a Banach space and ImI_{m} an open interval containing ss^{*}. Let F~:Im×XX\tilde{F}:I_{m}\times X\to X be a C2C^{2}-function with these properties:

(a) F~(s,0)=0\tilde{F}(s,0)=0 for sIms\in I_{m}.

(b) The partial derivatives D1F~D_{1}\tilde{F}, D2F~D_{2}\tilde{F}, D1D2F~D_{1}D_{2}\tilde{F}, and D22F~D_{2}^{2}\tilde{F} exist and are continuous.

(c) N(D2F~(s,0))N(D_{2}\tilde{F}(s^{*},0)) is one-dimensional, and R(D2F~(s,0))R(D_{2}\tilde{F}(s^{*},0)) has codimension 1.

(d) D1D2F~(s,0)(x0)R(D2F~(s,0))D_{1}D_{2}\tilde{F}(s^{*},0)(x_{0})\not\in R(D_{2}\tilde{F}(s^{*},0)), where N(D2F~(s,0))=Span({x0}).N(D_{2}\tilde{F}(s^{*},0))=\operatorname{Span}(\{x_{0}\}).

If ZZ is any complement of Span({x0})\operatorname{Span}(\{x_{0}\}) in XX, then there is a neighborhood UU of (s,0)(s^{*},0) in ×X\mathbb{R}\times X, an interval (a,a)(-a,a), and C1C^{1}-functions φ:(a,a)\varphi:(-a,a)\to\mathbb{R}, ψ:(a,a)Z\psi:(-a,a)\to Z such that φ(0)=s\varphi(0)=s^{*}, ψ(0)=0\psi(0)=0 and

F~1({0})U={(φ(α),αx0+αψ(α)):|α|<a}{(s,0):(s,0)U}.\tilde{F}^{-1}(\{0\})\cap U=\{(\varphi(\alpha),\alpha x_{0}+\alpha\psi(\alpha)):|\alpha|<a\}\cup\{(s,0):(s,0)\in U\}.
Remark 2.11.

Theorem 2.10 is not exactly as originally stated in [4, Theorem 1.7]. Our statement is convenient for our purposes and is an easy consequence of the original result, as now we explain.

First, we scale the parameter sIms\in I_{m} to t(1,1)t\in(-1,1). Choose ε>0\varepsilon>0 such that (sε,s+ε)Im(s^{*}-\varepsilon,s^{*}+\varepsilon)\subseteq I_{m}. We apply [4, Theorem 1.7] to F:(1,1)×XXF:(-1,1)\times X\to X defined by F(t,x)=F~(s+εt,x)F(t,x)=\tilde{F}(s^{*}+\varepsilon t,x). Hypothesis (b) in [4, Theorem 1.7] does not require Fxx=D22FF_{xx}=D_{2}^{2}F to be continuous, and the conclusion is that φ\varphi and ψ\psi are continuous. In a remark after the theorem they state that if FxxF_{xx} is continuous, then φ\varphi and ψ\psi are C1C^{1}, and this is the result stated in Theorem 2.10.

We now present our main result, which applies the BSE to nested invariant subspaces. Recall from Proposition 2.8 that the mother branch includes {(s,bm(s)):sIm}\{(s,b_{m}(s)):s\in I_{m}\}, with bm(s)Wmb_{m}(s)\in W_{m} and F(s,bm(s))=0F(s,b_{m}(s))=0 for sIms\in I_{m}, and bm(s)=xb_{m}(s^{*})=x^{*}.

Theorem 2.12 (Bifurcation Lemma for Invariant Subspaces).

Assume that the hypotheses of Proposition 2.8 hold. Let bmb_{m} and CmC_{m} be the mother branch function and mother branch. Assume WmWdXW_{m}\subsetneq W_{d}\subseteq X be nested FF-invariant subspaces. Let Fd:I×WdWdF_{d}:I\times W_{d}\to W_{d} be the restriction of FF. Assume J:ImL(Wd)J:I_{m}\to L(W_{d}), J(s):=D2Fd(s,bm(s))J(s):=D_{2}F_{d}(s,b_{m}(s)) satisfies the following conditions:

(a) N(J(s))N(J(s^{*})) is one-dimensional, and R(J(s))R(J(s^{*})) has codimension one.

(b) J(s)(x0)R(J(s))J^{\prime}(s^{*})(x_{0})\notin R(J(s^{*})), where N(J(s))=Span({x0})N(J(s^{*}))=\operatorname{Span}(\{x_{0}\}).

If ZZ is any complement of Span({x0})\operatorname{Span}(\{x_{0}\}) in WdW_{d}, then there is a neighborhood UU of (s,x)(s^{*},x^{*}) in I×XI\times X and C1C^{1}-functions φ:(a,a)\varphi:(-a,a)\to\mathbb{R} and ψ:(a,a)Wd\psi:(-a,a)\to W_{d} with φ(0)=s\varphi(0)=s^{*} and ψ(0)=0\psi(0)=0 such that

F1({0})(I×Wd)U=CmCd,F^{-1}(\{0\})\cap(I\times W_{d})\cap U=C_{m}\cup C_{d},

where

Cd:={(φ(α),bm(φ(α))+αx0+αψ(α)):|α|<a}C_{d}:=\{(\varphi(\alpha),b_{m}(\varphi(\alpha))+\alpha x_{0}+\alpha\psi(\alpha)):|\alpha|<a\}

is the so-called daughter branch. Furthermore, CmCd={(s,x)}C_{m}\cap C_{d}=\{(s^{*},x^{*})\}. That is, there are exactly two branches in (I×Wd)U(I\times W_{d})\cap U that contain (s,x)(s^{*},x^{*}); the mother branch CmC_{m} and the daughter branch CdC_{d}.

Proof.

Since WdW_{d} is FF-invariant, the restriction Fd:I×WdWdF_{d}:I\times W_{d}\to W_{d} is well-defined. We will show that the hypotheses of Theorem 2.10 apply to the function F~:Im×WdWd\tilde{F}:I_{m}\times W_{d}\to W_{d} defined by

F~(s,x)=Fd(s,bm(s)+x).\tilde{F}(s,x)=F_{d}(s,b_{m}(s)+x).

Condition (a) holds, since F~(s,0)=Fd(s,bm(s))=0\tilde{F}(s,0)=F_{d}(s,b_{m}(s))=0, by the definition of the mother branch function bmb_{m}. The restriction to FdF_{d} is allowed since bm(s)WmWdb_{m}(s)\in W_{m}\subseteq W_{d}. Condition (b) holds: F~\tilde{F} is C2C^{2} since FF is C2C^{2} and bmb_{m} is C2C^{2}. Conditions (c) and (d) require the computation of D2F~(s,0)D_{2}\tilde{F}(s^{*},0) and D1D2F~(s,0)D_{1}D_{2}\tilde{F}(s^{*},0). The Jacobian is

D2F~(s,x)=D2Fd(s,bm(s)+x),D_{2}\tilde{F}(s,x)=D_{2}F_{d}(s,b_{m}(s)+x),

and the stability of the solution at any ss on the mother branch is determined by

(2) J(s):=D2Fd(s,bm(s))=D2F~(s,0).J(s):=D_{2}F_{d}(s,b_{m}(s))=D_{2}\tilde{F}(s,0).

While J(s)J(s) is defined in terms of the first expression in Equation (2), we will use the second form of the expression for the remainder of this proof. Condition (a) of Theorem 2.12 concerns

J(s)=D2F~(s,0)J(s^{*})=D_{2}\tilde{F}(s^{*},0)

and is equivalent to Condition (c) of Theorem 2.10. Condition (b) of Theorem 2.12 concerns

J(s)=D1D2F~(s,0),J^{\prime}(s)=D_{1}D_{2}\tilde{F}(s,0),

and is equivalent to Condition (d) of Theorem 2.10.

Thus, Theorem 2.10 holds for the function F~\tilde{F}, and for the functions φ\varphi and ψ\psi defined in that theorem,

F~(φ(α),αx0+αψ(α))=0,\tilde{F}(\varphi(\alpha),\alpha x_{0}+\alpha\psi(\alpha))=0,

for |α|<a|\alpha|<a. By the definition of F~\tilde{F} in terms of FdF_{d},

Fd(φ(α),bm(φ(α))+αx0+αψ(α))=0,F_{d}(\varphi(\alpha),b_{m}(\varphi(\alpha))+\alpha x_{0}+\alpha\psi(\alpha))=0,

for |α|<a|\alpha|<a, and the daughter branch is given parametrically as desired. Theorem 2.10 ensures that the mother and daughter branch are the only local branches in I×WdI\times W_{d}. Finally, the point on the daughter branch CdC_{d} when α=0\alpha=0 is (φ(0),bm(φ(0)))=(s,x)(\varphi(0),b_{m}(\varphi(0)))=(s^{*},x^{*}). There is no other point on the daughter branch in I×WmI\times W_{m}, since bm(s)Wmb_{m}(s)\in W_{m}, and x0Wmx_{0}\not\in W_{m}, and ψ(α)\psi(\alpha) is in ZZ. Thus, the only point of intersection of CmC_{m} and CdC_{d} is (s,x)(s^{*},x^{*}). ∎

Remark 2.13.

We give several observations.

  1. (1)

    Condition (b) in Theorem 2.12 typically holds in applications, but it is difficult to check when the mother branch function is not known explicitly. This nondegeneracy condition implies that a simple eigenvalue of J(s)J(s) crosses 0 with nonzero speed at s=ss=s^{*}. This leads to an informal statement of the BLIS:

    Suppose WmWdW_{m}\subsetneq W_{d} are invariant subspaces of F:×XXF:\mathbb{R}\times X\to X, with F(s,x)=0F(s^{*},x^{*})=0 for xWmx^{*}\in W_{m}. Let K=N(D2F(s,x))K=N(D_{2}F(s^{*},x^{*})). If

    KWm={0} and dim(KWd)=1,K\cap W_{m}=\{0\}\text{ and }\dim(K\cap W_{d})=1,

    then generically F(s,x)=0F(s,x)=0 has exactly two solution branches in ×Wd\mathbb{R}\times W_{d} that contain (s,x)(s^{*},x^{*}); the mother branch which is in ×Wm\mathbb{R}\times W_{m}, and the daughter branch.

    See [11, pg. 17] for a discussion of the meaning of “generically,” which is rather subtle. In summary, generic behavior is typical.

  2. (2)

    Every EBL bifurcation is a BLIS bifurcation. That is, our Theorem 2.12 implies the results usually obtained by an equivariant Lyapunov-Schmidt reduction followed by an application of the EBL [3, 12, 25]:

    Assume that a compact Lie group Γ\Gamma acts on a Banach space XX, that F:I×XXF:I\times X\to X is Γ\Gamma-equivariant, and that Σd<ΣmΓ\Sigma_{d}<\Sigma_{m}\leq\Gamma. Let xFix(Σm)x^{*}\in\operatorname{Fix}(\Sigma_{m}) satisfy F(s,x)=0F(s^{*},x^{*})=0 and let K=N(D2F(s,x))K=N(D_{2}F(s^{*},x^{*})). If

    KFix(Σm)={0} and dim(KFix(Σd))=1,K\cap\operatorname{Fix}(\Sigma_{m})=\{0\}\text{ and }\dim(K\cap\operatorname{Fix}(\Sigma_{d}))=1,

    then generically a branch of solutions in I×Fix(Σd)I\times\operatorname{Fix}(\Sigma_{d}) bifurcates from the mother branch, which is in I×Fix(Σm)I\times\operatorname{Fix}(\Sigma_{m}), at (s,x)(s^{*},x^{*}).

  3. (3)

    Our Theorem 2.12 is a generalization of the Synchrony Branching Lemma for regular coupled cell networks. See [9, Theorem 6.3], [5, Theorem 8.2] and [11, Theorem 18.15]. This theorem was generalized in [24]. We state this theorem, indicating the mother and daughter subspaces:

    Let XX be the finite dimensional phase space of a network with the admissible map F:×XXF:\mathbb{R}\times X\to X, defined in [5]. Let Wm=ΔW_{m}=\Delta be the space of fully synchronous states, and let Wd=ΔW_{d}=\Delta_{\bowtie} be the synchrony subspace for some balanced coloring \bowtie. Let xΔx^{*}\in\Delta, ss^{*}\in\mathbb{R}, and let K=N(D2F(s,x))K=N(D_{2}F(s^{*},x^{*})). If

    KΔ={0} and dim(KΔ)=1,K\cap\Delta=\{0\}\text{ and }\dim(K\cap\Delta_{\bowtie})=1,

    then generically a unique branch of equilibria with synchrony pattern {\bowtie} bifurcates from (s,x)(s^{*},x^{*}).

  4. (4)

    There might be many other branches bifurcating from (s,x)(s^{*},x^{*}). When we restrict to branches in I×WdI\times W_{d}, however, there are exactly two branches (the mother and daughter branches).

  5. (5)

    Note that

    N(J(s)):=N(D2Fd(s,x))=N(D2F(s,x))Wd.N(J(s^{*})):=N(D_{2}F_{d}(s^{*},x^{*}))=N(D_{2}F(s^{*},x^{*}))\cap W_{d}.

    As in Part (1) of this remark, let K=N(D2F(s,x))K=N(D_{2}F(s^{*},x^{*})). One of the hypotheses in Proposition 2.8 is equivalent to KWm={0}K\cap W_{m}=\{0\}. Condition (a) of Theorem 2.12 implies that dim(KWd)=1\dim(K\cap W_{d})=1.

  6. (6)

    If WdW_{d} is finite dimensional, then the two parts of Condition (a) of Theorem 2.12 are equivalent. That is, we only need to check that N(J(s))N(J(s^{*})) is one-dimensional.

  7. (7)

    Rather than having one daughter branch that crosses the mother branch at the bifurcation point, we often consider two daughter branches Cd+C_{d}^{+} and CdC_{d}^{-}, born at the bifurcation, defined by

    Cd±:={(φ(α),bm(φ(α)+αx0+αψ(α)):0<±α<a}.C_{d}^{\pm}:=\{(\varphi(\alpha),b_{m}(\varphi(\alpha)+\alpha x_{0}+\alpha\psi(\alpha)):0<\pm\alpha<a\}.

    Note that 0<α<a0<\alpha<a for the parametric curve Cd+C_{d}^{+} while a<α<0-a<\alpha<0 for CdC_{d}^{-}. Neither of these branches include (s,x)(s^{*},x^{*}), and

    CmCd=Cm˙Cd+˙Cd.C_{m}\cup C_{d}=C_{m}\dot{\cup}C_{d}^{+}\dot{\cup}C_{d}^{-}.

    The MATLAB® code, described in Section 4 and freely available at [18], follows the branches Cd+C_{d}^{+} and CdC_{d}^{-} separately, starting from their birth at (s,x)(s^{*},x^{*}). At a pitchfork bifurcation where the two daughter branches are conjugate by a symmetry of the system, the code only follows Cd+C_{d}^{+}.

3. Applications of the BLIS

We consider several examples of bifurcations and classify them as BSE, EBL, or BLIS, as indicated in Figure 1.

3.1. One-Dimensional Examples

We consider some applications of BSE, EBL, and BLIS in the case where X=X=\mathbb{R}. We will write examples of ODES, x˙=F(s,x)\dot{x}=F(s,x), and look for the equilibrium solutions which satisfy F(s,x)=0F(s,x)=0, for functions F:F:\mathbb{R}\to\mathbb{R}. In such cases, J(s)J(s) in the statement of Theorem 2.12 is a 1×11\times 1 matrix.

Example 3.1.

The ODE

x˙=F(s,x):=sxx3\dot{x}=F(s,x):=sx-x^{3}

has a pitchfork bifurcation at (s,x)=(0,0)(s,x)=(0,0). In this case, Wm={0}Wd=W_{m}=\{0\}\subseteq W_{d}=\mathbb{R}. Factoring FF shows that the mother branch has the equation x=0x=0, and the daughter branch has the equation s=x2s=x^{2}.

The mother branch function is bm(s)=0b_{m}(s)=0 for all ss, and J(s)=s[s]1×1J(s)=s\equiv[s]\in\mathbb{R}^{1\times 1}. The bifurcation point is s=0s^{*}=0, and J(0)=0J(0)=0, and J(0)=1J^{\prime}(0)=1. The system has the symmetry F(s,x)=F(s,x)F(s,-x)=-F(s,x), so both WmW_{m} and WdW_{d} are fixed point subspaces. The bifurcation is BSE, EBL, and BLIS. See Figure 2(a).

Example 3.2.

The ODE

x˙=F(s,x):=sxx2\dot{x}=F(s,x):=sx-x^{2}

has a so-called transcritical bifurcation at (s,x)=(0,0)(s,x)=(0,0). In this case, Wm={0}Wd=W_{m}=\{0\}\subseteq W_{d}=\mathbb{R}. Factoring FF shows that the mother branch has the equation x=0x=0, and the daughter branch has the equation s=xs=x.

The mother branch function is bm(s)=0b_{m}(s)=0 for all ss, and J(s)=[s]1×1J(s)=[s]\in\mathbb{R}^{1\times 1}. The bifurcation point is s=0s^{*}=0, and J(0)=[0]J(0)=[0], and J(0)=[1]J^{\prime}(0)=[1]. The trivial subspace is FF-invariant since F(s,0)=0F(s,0)=0 for all ss, The bifurcation is BSE and BLIS. See Figure 2(b).

Example 3.3.

The ODE

x˙=F(s,x):=s2x2\dot{x}=F(s,x):=s^{2}-x^{2}

has a different kind of transcritical bifurcation at (s,x)=(0,0)(s,x)=(0,0). In this case, the only FF-invariant subspace is \mathbb{R}, so this is not a BLIS bifurcation. The function FF has no symmetry, so this is not an EBL bifurcation. Factoring FF shows that there are two branches, x=sx=s and s=xs=-x. Arbitrarily choose one of these branches, x=bm(s):=sx=b_{m}(s):=-s and define F~(s,x)=F(s,xbm(s))=s2(x+s)2=2sxx2\tilde{F}(s,x)=F(s,x-b_{m}(s))=s^{2}-(x+s)^{2}=2sx-x^{2}. The BSE Theorem 2.10 and BLIS Theorem 2.12 both apply to F~\tilde{F} with s=0s^{*}=0. The relevant calculations are that D2F~(s,x)=2s2xD_{2}\tilde{F}(s,x)=2s-2x, so D2F~(0,0)=0D_{2}\tilde{F}(0,0)=0 and D1D2F(s,x)=2D_{1}D_{2}F(s,x)=2. The critical eigenvector is x0=1x_{0}=1, so Span({x0})=\operatorname{Span}(\{x_{0}\})=\mathbb{R}. It is clear from factoring F~\tilde{F} that the two branches of F~(s,x)=0\tilde{F}(s,x)=0 are x=2sx=2s and x=0x=0, so

F~1({0})={(α/2,α):α}{(s,0):s}.\tilde{F}^{-1}(\{0\})=\{(\alpha/2,\alpha):\alpha\in\mathbb{R}\}\cup\{(s,0):s\in\mathbb{R}\}.

Theorems 2.10 and 2.12 hold with ϕ(α)=α/2\phi(\alpha)=\alpha/2.

Theorem 1 of [4] describes how the original function FF, which a branch of nontrivial solutions, has a bifurcation. Thus we say that FF has a BSE bifurcation, as indicated in Figure 1.

Although Theorem 2.12 applies to the reduced function F~\tilde{F}, it does not apply directly to FF. Since the original function FF has no nested invariant subspaces, we do not include this example in the BLIS region of Figure 1.

Note that this example does not occur generically in one-parameter families. If a small term ε\varepsilon is added, then the system becomes x˙=F(s,ε,x):=s2+εx2\dot{x}=F(s,\varepsilon,x):=s^{2}+\varepsilon-x^{2}, which has a qualitatively different bifurcation diagram of xx vs. ss for fixed ε>0\varepsilon>0, ε=0\varepsilon=0, or ε<0\varepsilon<0. In contrast, the transcritical bifurcation in Example 3.2 is generic for the class of systems with F(s,0)=0F(s,0)=0, which is a natural constraint for some models. See Figure 2(c).

Example 3.4.

The ODE

x˙=F(s,x):=sx2\dot{x}=F(s,x):=s-x^{2}

has the equilibrium solutions

F1({0})={(α2,α):α}.F^{-1}(\{0\})=\{(\alpha^{2},\alpha):\alpha\in\mathbb{R}\}.

The point (s,x)=(0,0)(s,x)=(0,0) is a fold point, We do not call this a saddle-node bifurcation, following our Definition 2.7. Since there is only one solution branch and no branching of solutions, neither the BSE nor the BLIS apply. See Figure 2(d).

(a)(b)(c)(d)(e)
Figure 2. The xx vs. ss bifurcation diagrams of the ODEs featured in Examples 3.1(a), 3.2(b), 3.3(c), 3.4(d), and 3.5(e). In each case the dot is at (s,x)=(0,0)(s,x)=(0,0), and the solid lines represent a branch of stable equilibrium solutions to x˙=F(s,x)\dot{x}=F(s,x). The dashed lines represent a branch of unstable equilibrium solutions.
Example 3.5.

Consider the ODE

x˙=F(s,x):=s2xx3=x(s+x)(sx).\dot{x}=F(s,x):=s^{2}x-x^{3}=x(s+x)(s-x).

The mother branch function is bm(s)=0b_{m}(s)=0, and the stability of the trivial branch is determined by J(s)=s2J(s)=s^{2}. Condition (a) of Theorem 2.12 holds, with (s,x)=(0,0)(s^{*},x^{*})=(0,0), Wm={0}W_{m}=\{0\}, Wd=W_{d}=\mathbb{R}, and the critical eigenvector x0=1x_{0}=1. However the nondegeneracy Condition (b) of Theorem 2.12 does not hold, since J(s)=2sJ^{\prime}(s)=2s and thus J(0)(1)=0R(J(0))J^{\prime}(0)(1)=0\in R(J(0)). Similarly, the BSE does not hold for this example, since Condition (d) of Theorem 2.10 is false. The function FF does have symmetry, but the EBL does not hold. This nongeneric example is one of the exceptions referred to in Remark 2.13(1). The BLIS says that generically there are exactly two solution branches in ×X\mathbb{R}\times X that contain (s,x)(s^{*},x^{*}). Similar examples of F:×XXF:\mathbb{R}\times X\to X for which Condition (a) of the BLIS holds, but Condition (b) does not hold, can be manufactured for any XX. See Figure 2(e).

3.2. Network Examples

We apply the BLIS to systems of networks of nn coupled cells defined by x˙=F(s,x)\dot{x}=F(s,x) for F:×(k)n(k)nF:\mathbb{R}\times(\mathbb{R}^{k})^{n}\to(\mathbb{R}^{k})^{n}. The ODE for each cell with phase space xikx_{i}\in\mathbb{R}^{k} is

(3) x˙i=F(s,x)i:=f(s,xi)+Hj=1nMi,jxj=f(s,xi)+H(Mx)i,{\dot{x}}_{i}=F(s,x)_{i}:=f(s,x_{i})+H\sum_{j=1}^{n}M_{i,j}x_{j}=f(s,x_{i})+H(Mx)_{i},

where f:×kkf:\mathbb{R}\times\mathbb{R}^{k}\to\mathbb{R}^{k} describes the internal dynamics of each identical cell, Hk×kH\in\mathbb{R}^{k\times k} is the coupling matrix, and Mn×nM\in\mathbb{R}^{n\times n} is the weighted adjacency matrix of the network. Note that we multiply a real matrix Mn×nM\in\mathbb{R}^{n\times n} with a matrix x(k)n×1x\in(\mathbb{R}^{k})^{n\times 1} whose entries come from k\mathbb{R}^{k}, resulting in a matrix Mx(k)n×1Mx\in(\mathbb{R}^{k})^{n\times 1}, which we identify with (k)n(\mathbb{R}^{k})^{n}.

In order to understand the BLIS bifurcations, we first describe the subspaces that are invariant under System (3) for a fixed MM with any ff and HH. These invariant subspaces are related to the MM-invariant polydiagonal subspaces of n\mathbb{R}^{n}, as described in [21]. Here we show that the dynamics on one of these invariant subspaces is also described by System (3) but with dd coupled cells, where the n×nn\times n matrix MM is replaced by a smaller d×dd\times d matrix. We can think of any matrix MM as the adjacency matrix of a weighted digraph with nn vertices, and then the new d×dd\times d matrix is the weighted adjacency matrix of the weighted quotient digraph with dd vertices.

We use the definition of polydiagonal subspaces found in [21], which includes both synchrony and anti-synchrony subspaces. This development builds on the work of [1, 11], but the terminology is different. Here we provide a characterization of a nontrivial polydiagonal subspace as the column space of a matrix with special properties. This new characterization lends itself well to numerical computations, and is used in our MATLAB® program [18].

Definition 3.6.

For 0<dn0<d\leq n, a polydiagonal basis matrix is a matrix B{1,0,1}n×dB\in\{1,0,-1\}^{n\times d} that satisfies the following three conditions:

  1. (1)

    rank(B)=d\text{rank}\,(B)=d;

  2. (2)

    every row of BB has at most one non-zero entry;

  3. (3)

    BTB^{T} is in reduced row-echelon form.

If, in addition, every row of BB has an entry of 1, then BB is called a synchrony basis matrix. Otherwise, BB is called an anti-synchrony basis matrix.

Thus, a synchrony basis matrix only has elements 0 and 1, and every row has a single 1. An anti-synchrony basis matrix has a row of all 0’s or a row with a single 1-1.

Proposition 3.7.

A nontrivial polydiagonal subspace of n\mathbb{R}^{n} is the column space of a unique polydiagonal matrix. Furthermore, Col(B)\operatorname{Col}(B) is a synchrony/anti-synchrony subspace if and only if BB is a synchrony/anti-synchrony basis matrix.

Remark 3.8.

The proof of this proposition is straightforward but tedious. The connection with the definition of a polydiagonal subspace in [1, 21] is that a polydiagonal subspace and Col(B)\operatorname{Col}(B) can both be defined by equations of the form xi=xjx_{i}=x_{j} or xi=xjx_{i}=-x_{j} (including xi=0x_{i}=0 when i=ji=j). Any set of equations defining an anti-synchrony subspace has at least one equation of the form xi=xjx_{i}=-x_{j}. Conditions (1) and (3) in Definition 3.6 ensure that Col(B1)=Col(B2)\operatorname{Col}(B_{1})=\operatorname{Col}(B_{2}) implies B1=B2B_{1}=B_{2}.

Note that the trivial subspace {0}n\{0\}\subseteq\mathbb{R}^{n} is an anti-synchrony subspace according to the definition in [21], but it is not the column space of any non-empty matrix.

Example 3.9.

The matrix B:=[101010]TB:=\left[\begin{smallmatrix}1&0&1\\ 0&1&0\end{smallmatrix}\right]^{T} is the synchrony basis matrix for {(a,b,a):a,b}3\{(a,b,a):a,b\in\mathbb{R}\}\subseteq\mathbb{R}^{3}. The corresponding tagged partition, as defined in [21], is 𝒫={{1,3},{2}}\mathcal{P}=\{\{1,3\},\{2\}\} with the empty partial involution.

The matrix B:=[10010100]TB:=\left[\begin{smallmatrix}1&0&0&-1\\ 0&1&0&0\end{smallmatrix}\right]^{T} is the anti-synchrony basis matrix for {(a,b,0,a):a,b}4\{(a,b,0,-a):a,b\in\mathbb{R}\}\subseteq\mathbb{R}^{4}. The corresponding tagged partition is 𝒫={{1},{2},{3},{4}}\mathcal{P}=\{\{1\},\{2\},\{3\},\{4\}\} with {1}={4}\{1\}^{*}=\{4\} and {3}={3}\{3\}^{*}=\{3\}. Note that {2}\{2\} is not in the domain of the partial involution.

Let Bn×dB\in\mathbb{R}^{n\times d} be a polydiagonal basis matrix. Then Col(B)\operatorname{Col}(B) is MM-invariant if and only if Col(MB)Col(B)\operatorname{Col}(MB)\subseteq\operatorname{Col}(B). See [17] for an efficient way to find the MM-invariant synchrony subspaces of n\mathbb{R}^{n}. Note that the columns of BB are orthogonal due to Condition (2), and BTBB^{T}B is a diagonal d×dd\times d matrix, with (BTB),(B^{T}B)_{\ell,\ell} equal to the number of nonzero components in the \ellth column of BB. Condition (1) implies that (BTB),>0(B^{T}B)_{\ell,\ell}>0. The Moore-Penrose inverse (pseudoinverse)

B+=(BTB)1BTd×nB^{+}=(B^{T}B)^{-1}B^{T}\in\mathbb{R}^{d\times n}

of BB is easily computed since BTBB^{T}B is diagonal and nonsingular. The pseudoinverse satisfies B+B=IdB^{+}B=I_{d} and BB+BB^{+} is the projection of n\mathbb{R}^{n} onto Col(B)\operatorname{Col}(B).

Example 3.10.

Let B=[10010100]TB=\left[\begin{smallmatrix}1&0&0&-1\\ 0&1&0&0\end{smallmatrix}\right]^{T}. Then BTB=[2001]B^{T}B=\left[\begin{smallmatrix}2&0\\ 0&1\end{smallmatrix}\right], so

B+=[2001]1[10010100]=[0.5000.50100]B^{+}=\begin{bmatrix}2&0\\ 0&1\end{bmatrix}^{-1}\begin{bmatrix}1&0&0&-1\\ 0&1&0&0\end{bmatrix}=\begin{bmatrix}0.5&0&0&-0.5\\ 0&1&0&0\end{bmatrix}

is the pseudoinverse of BB.

We now consider the dynamically invariant subspaces of System (3), which are precisely the FF-invariant subspaces. Proposition 4.1 of [21] describes these subspaces. If MM and H0H\neq 0 are fixed, then the nontrivial subspaces of (k)n(\mathbb{R}^{k})^{n} that are FF-invariant for all odd ff are precisely the subspaces of the form

WB:={By:y(k)d}(k)nW_{B}:=\{By:y\in(\mathbb{R}^{k})^{d}\}\subseteq(\mathbb{R}^{k})^{n}

for some polydiagonal basis matrix Bn×dB\in\mathbb{R}^{n\times d} such that Col(B)\operatorname{Col}(B) is MM-invariant. The trivial subspace {0}(k)n\{0\}\subseteq(\mathbb{R}^{k})^{n} FF-invariant whenever f(0)=0f(0)=0. Furthermore, if ff is not odd and BB is a synchrony basis matrix, then WBW_{B} is FF-invariant.

Note that By(k)d×1(k)dBy\in(\mathbb{R}^{k})^{d\times 1}\equiv(\mathbb{R}^{k})^{d} is a product similar to MxMx in System (3). If k=1k=1, then WB=Col(B)W_{B}=\operatorname{Col}(B). If k>1k>1, then WB=kCol(B)W_{B}=\mathbb{R}^{k}\otimes\operatorname{Col}(B), which is the tensor product of k\mathbb{R}^{k} with the polydiagonal subspace Col(B)\operatorname{Col}(B) [21].

Example 3.11.

Let B=[10010100]TB=\left[\begin{smallmatrix}1&0&0&-1\\ 0&1&0&0\end{smallmatrix}\right]^{T} and k=3k=3, so that y=(y1,y2)(3)2y=(y_{1},y_{2})\in(\mathbb{R}^{3})^{2}. We find that

By=[10010010][y1y2]=[y1y20y1](y1,y2,0,y1),By=\left[\begin{matrix}1&0\\ 0&1\\ 0&0\\ -1&0\end{matrix}\right]\left[\begin{matrix}y_{1}\\ y_{2}\end{matrix}\right]=\left[\begin{matrix}y_{1}\\ y_{2}\\ 0\\ -y_{1}\end{matrix}\right]\equiv(y_{1},y_{2},0,-y_{1}),

so WB={(y1,y2,0,y1):y1,y23}(3)4W_{B}=\{(y_{1},y_{2},0,-y_{1}):y_{1},y_{2}\in\mathbb{R}^{3}\}\subseteq(\mathbb{R}^{3})^{4}. A typical element of WBW_{B} is (y1,y2,0,y1)(y_{1},y_{2},0,-y_{1}), which we can abbreviate to (a,b,0,a)(a,b,0,-a), as in Figure 3.

We now describe how to restrict the System (3) to an invariant subspace WBW_{B}. To simplify the exposition we first consider a network of uncoupled cells with no parameter.

Proposition 3.12.

Let Bn×dB\in\mathbb{R}^{n\times d} be a polydiagonal basis matrix, f:kkf:\mathbb{R}^{k}\to\mathbb{R}^{k}, and F:(k)n(k)nF:(\mathbb{R}^{k})^{n}\to(\mathbb{R}^{k})^{n} defined by F(x)i:=f(xi)F(x)_{i}:=f(x_{i}). If ff is odd or BB is a synchrony basis matrix, then FB:(k)d(k)dF_{B}:(\mathbb{R}^{k})^{d}\to(\mathbb{R}^{k})^{d} defined by FB(y):=B+F(By)F_{B}(y):=B^{+}F(By) satisfies

FB(y)=f(y).F_{B}(y)_{\ell}=f(y_{\ell}).
Proof.

Let

N:={i() Bi,0}N:=\{i\mid(\exists\ell)\text{ }B_{i,\ell}\neq 0\}

be the set of indices ii such that the iith row of BB has a nonzero element. Define g:N{1,,d}g:N\to\{1,\ldots,d\} and σ:N{1,1}\sigma:N\to\{1,-1\} such that

Bi,={σ(i)if =g(i)0otherwise.B_{i,\ell}=\begin{cases}\sigma(i)&\text{if }\ell=g(i)\\ 0&\text{otherwise.}\end{cases}

If iNi\not\in N, then (By)i=0(By)_{i}=0. If iNi\in N, then (By)i==1nBi,y=σ(i)yg(i)(By)_{i}=\sum_{\ell=1}^{n}B_{i,\ell}y_{\ell}=\sigma(i)y_{g(i)} and F(By)i=f((By)i)=f(σ(i)yg(i))F(By)_{i}=f((By)_{i})=f(\sigma(i)y_{g(i)}).

If ff is odd, then F(By)i=f(0)=0F(By)_{i}=f(0)=0 for iNi\not\in N and F(By)i=f(σ(i)yg(i))=σ(i)f(yg(i))F(By)_{i}=f(\sigma(i)y_{g(i)})=\sigma(i)f(y_{g(i)}) for iNi\in N. If BB is a synchrony basis matrix, then σ(i)=1\sigma(i)=1 for all iN={1,,n}i\in N=\{1,\ldots,n\}. In either case,

F(By)i={σ(i)f(yg(i))if iN0if iN.F(By)_{i}=\begin{cases}\sigma(i)f(y_{g(i)})&\text{if }i\in N\\ 0&\text{if }i\not\in N.\end{cases}

Let h():=(BTB),h(\ell):=(B^{T}B)_{\ell,\ell} be the number of nonzero elements in the \ellth column of BB. Note that h()h(\ell) is the size of the set g1()g^{-1}(\ell). The pseudoinverse of BB has components

B,i+={σ(i)h()if =g(i)0otherwise.B^{+}_{\ell,i}=\begin{cases}\frac{\sigma(i)}{h(\ell)}&\text{if }\ell=g(i)\\ 0&\text{otherwise.}\end{cases}

Then

FB(y)=i=1nB,i+F(By)i=ig1()σ(i)h()σ(i)f(yg(i))=h()f(y)h()=f(y)F_{B}(y)_{\ell}=\sum_{i=1}^{n}B^{+}_{\ell,i}F(By)_{i}=\sum_{i\in g^{-1}(\ell)}\frac{\sigma(i)}{h(\ell)}\sigma(i)f(y_{g(i)})=h(\ell)\frac{f(y_{\ell})}{h(\ell)}=f(y_{\ell})

since the sum has h()h(\ell) nonzero terms, all equal to f(y)/h()f(y_{\ell})/h(\ell). ∎

Remark 3.13.

Proposition 3.12 is the motivation for the definition of a polydiagonal basis matrix. Consider B=[2 0]TB=[2\ 0]^{T}, which is not a polydiagonal basis matrix since B{1,0,1}n×dB\not\in\{1,0,-1\}^{n\times d} even though Col(B)\operatorname{Col}(B) is a polydiagonal subspace which is FF-invariant for the uncoupled system in Proposition 3.12 if f(0)=0f(0)=0. Since FB(y)1=12f(2y1)F_{B}(y)_{1}=\frac{1}{2}f(2y_{1}) and in general 12f(2y1)f(y1)\frac{1}{2}f(2y_{1})\not=f(y_{1}), the single reduced cell y˙1=FB(y)1\dot{y}_{1}=F_{B}(y)_{1} may not have the internal dynamics y˙1=f(y1)\dot{y}_{1}=f(y_{1}).

We now use Proposition 3.12 to write FBF_{B} in terms of yky\in\mathbb{R}^{k}, including the parameter ss.

Proposition 3.14.

Let Bn×dB\in\mathbb{R}^{n\times d} be a polydiagonal basis matrix, Col(B)\operatorname{Col}(B) be an MM-invariant polydiagonal subspace, Hk×kH\in\mathbb{R}^{k\times k}, f:×kkf:\mathbb{R}\times\mathbb{R}^{k}\to\mathbb{R}^{k}, and F:×(k)n(k)nF:\mathbb{R}\times(\mathbb{R}^{k})^{n}\to(\mathbb{R}^{k})^{n} be defined by

(4) F(s,x)i:=f(s,xi)+Hj=1nMi,jxj.F(s,x)_{i}:=f(s,x_{i})+H\sum_{j=1}^{n}M_{i,j}x_{j}.

Assume that ff is odd or BB is a synchrony basis matrix. The function FB:×(k)d(k)dF_{B}:\mathbb{R}\times(\mathbb{R}^{k})^{d}\to(\mathbb{R}^{k})^{d} defined by FB(s,y)=B+F(s,By)F_{B}(s,y)=B^{+}F(s,By) satisfies

(5) FB(s,y)=f(s,y)+Hm=1d(B+MB),mym.F_{B}(s,y)_{\ell}=f(s,y_{\ell})+H\sum_{m=1}^{d}(B^{+}MB)_{\ell,m}y_{m}.
Proof.

The internal dynamics ff is handled in Proposition 3.12, and the inclusion of the parameter ss in both ff and FF has no effect on the result. Recall that x=Byx=By means that xj=(By)j=m=1dBj,mymx_{j}=(By)_{j}=\sum_{m=1}^{d}B_{j,m}y_{m}. Thus,

FB(s,y)\displaystyle F_{B}(s,y)_{\ell} =i=1nB,i+F(s,By)i\displaystyle=\sum_{i=1}^{n}B^{+}_{\ell,i}F(s,By)_{i}
=f(s,y)+i=1nB,i+(Hj=1nMi,j(By)j)\displaystyle=f(s,y_{\ell})+\sum_{i=1}^{n}B^{+}_{\ell,i}\left(H\sum_{j=1}^{n}M_{i,j}(By)_{j}\right)
=f(s,y)+Hi=1nj=1nm=1dB,i+Mi,jBj,mym\displaystyle=f(s,y_{\ell})+H\sum_{i=1}^{n}\sum_{j=1}^{n}\sum_{m=1}^{d}B^{+}_{\ell,i}M_{i,j}B_{j,m}y_{m}
=f(s,y)+Hm=1d(B+MB),mym.\displaystyle=f(s,y_{\ell})+H\sum_{m=1}^{d}(B^{+}MB)_{\ell,m}y_{m}.

Note the similarity of Equations (4) and (5). The internal dynamics f:×kkf:\mathbb{R}\times\mathbb{R}^{k}\to\mathbb{R}^{k} is the same in both equations, but Mn×nM\in\mathbb{R}^{n\times n} in Equation (4) is replaced by B+MBd×dB^{+}MB\in\mathbb{R}^{d\times d} in Equation (5). We interpret MM as the adjacency matrix of a weighted digraph with nn vertices, and B+MBB^{+}MB is the adjacency matrix of the weighted quotient digraph with dd vertices. See Figure 3. Note that a graph (which is by definition unweighted, undirected, and has no loops) can have a weighted quotient digraph which is weighted, directed, and has loops.

While the proofs of Propositions 3.12 and 3.14 are complicated, the result is obvious when the nn component equations of F(s,x)F(s,x) are restricted to a polydiagonal subspace WBW_{B}. It is clear that only d=dim(WB)d=\dim(W_{B}) equations are essential and the rest hold automatically, as demonstrated in the next example.

Example 3.15.

Let

M=[0111100110011110], and B=[10010010], so MB=[11200011]M=\begin{bmatrix}0&1&1&1\\ 1&0&0&-1\\ 1&0&0&1\\ 1&-1&1&0\end{bmatrix},\text{ and }B=\begin{bmatrix}1&0\\ 0&1\\ 0&0\\ -1&0\end{bmatrix},\text{ so }MB=\begin{bmatrix}-1&1\\ 2&0\\ 0&0\\ 1&-1\end{bmatrix}

Note that Col(B)=Col(MB)\operatorname{Col}(B)=\operatorname{Col}(MB), so WBW_{B} is MM-invariant. System (3) is x˙=F(s,x)\dot{x}=F(s,x), where

F(s,x)1\displaystyle F(s,x)_{1} =f(s,x1)+H(x2+x3+x4)\displaystyle=f(s,x_{1})+H(x_{2}+x_{3}+x_{4})
F(s,x)2\displaystyle F(s,x)_{2} =f(s,x2)+H(x1x4)\displaystyle=f(s,x_{2})+H(x_{1}-x_{4})
F(s,x)3\displaystyle F(s,x)_{3} =f(s,x3)+H(x1+x4)\displaystyle=f(s,x_{3})+H(x_{1}+x_{4})
F(s,x)4\displaystyle F(s,x)_{4} =f(s,x4)+H(x1x2+x3).\displaystyle=f(s,x_{4})+H(x_{1}-x_{2}+x_{3}).

Assume ff is odd. It follows that WBW_{B} is an FF-invariant subspace, and the ODE can be restricted to WBW_{B} by setting x3=0x_{3}=0, and x4=x1x_{4}=-x_{1}. We find that F(s,x)4=F(s,x)1F(s,x)_{4}=-F(s,x)_{1} and F(s,x)3=0F(s,x)_{3}=0, so only the first and second components are needed. Proposition 3.14 formalizes this procedure. The pseudoinverse of BB is computed in Example 3.10, which yields the adjacency matrix

B+MB=[1120]B^{+}MB=\begin{bmatrix}-1&1\\ 2&0\end{bmatrix}

of the weighted quotient digraph. System (3) restricted to WBW_{B} is y˙=FB(s,y)\dot{y}=F_{B}(s,y), where

FB(s,y)1\displaystyle F_{B}(s,y)_{1} =f(s,y1)+H(y1+y2)\displaystyle=f(s,y_{1})+H(-y_{1}+y_{2})
FB(s,y)2\displaystyle F_{B}(s,y)_{2} =f(s,y2)+H(2y1).\displaystyle=f(s,y_{2})+H(2y_{1}).

This example demonstrates that a weighted digraph with a symmetric adjacency matrix can give rise to a weighted quotient digraph with a non-symmetric adjacency matrix, as shown in Figure 3.

Refer to caption Refer to caption Refer to caption
(a) (b) (c)
Figure 3. For the matrices MM and BB defined in Example 3.15, this figure shows (a) the weighted digraph whose adjacency matrix is MM, (b) a typical element of the FF-invariant subspace WBW_{B}, and (c) the weighted quotient digraph with the adjacency matrix B+MBB^{+}MB. The weights of the arrows in the digraphs are all 1 unless indicated.

We have investigated System (3) for all of the connected graphs with 4 or fewer vertices, with MM set to the graph Laplacian matrix. Among these examples only the diamond graph has a bifurcation that is BLIS but neither BSE nor EBL.

Example 3.16.

Let GG be the 4-vertex diamond graph, shown in Figure 4. The graph Laplacian matrix of GG is

(6) L=[2101131101211113].L=\begin{bmatrix}\phantom{-}2&-1&\phantom{-}0&-1\\ -1&\phantom{-}3&-1&-1\\ \phantom{-}0&-1&\phantom{-}2&-1\\ -1&-1&-1&\phantom{-}3\end{bmatrix}.

The lattice of LL-invariant subspaces are shown in Figure 4.

Refer to caption
Refer to caption
Figure 4. The LL-invariant subspaces for the Laplacian matrix LL of the diamond graph GG. A typical element of each invariant subspace is shown on the left, with the exception of W0={0}4W_{0}=\{0\}\subseteq\mathbb{R}^{4}, W10=4W_{10}=\mathbb{R}^{4}, and a reflection of W5W_{5}. The lattice of group orbits of LL-invariant subspaces is shown on the right. The boxes indicate invariant subspaces that are not fixed point subspaces of the Aut(G)×2\operatorname{Aut}(G)\times\mathbb{Z}_{2} action. The dashed arrows connect invariant subspaces with the same point stabilizer.

Assume that the state space of each cell has dimension k=1k=1, and the internal dynamics is f(s,xi)=sxi+xi3f(s,x_{i})=sx_{i}+x_{i}^{3}. We set M=LM=L, and H=[1]1H=[-1]\equiv-1. Note that we identify the 1×11\times 1 matrix [1][-1] with a real number. Thus, System (3) becomes

(7) Fi(s,x)=sxi+xi3(Lx)i,i{1,2,3,4}.F_{i}(s,x)=sx_{i}+x_{i}^{3}-(Lx)_{i},\quad i\in\{1,2,3,4\}.

Thus we have a function F:×44F:\mathbb{R}\times\mathbb{R}^{4}\to\mathbb{R}^{4} whose component functions are Equation (7). We search for solutions to F(s,x)=0F(s,x)=0, and we want to understand the bifurcations. As proved in [21], the FF-invariant subspaces are the LL-invariant subspaces. These are shown in Figure 4 .

The eigenvalues of LL are λ=0\lambda=0 with eigenvector (1,1,1,1)(1,1,1,1), λ=2\lambda=2 with eigenvector (1,0,1,0)(1,0,-1,0), and λ=4\lambda=4 with an eigenspace spanned by (1,1,1,1)(1,-1,1,-1) and (0,1,0,1)(0,1,0,-1).

The vector field FF is Aut(G)×2\operatorname{Aut}(G)\times\mathbb{Z}_{2}-equivariant, where Aut(G)=(1 3),(2 4)2×2\operatorname{Aut}(G)=\langle(1\,3),(2\,4)\rangle\cong\mathbb{Z}_{2}\times\mathbb{Z}_{2} is the symmetry group of the graph GG. The extra 2\mathbb{Z}_{2} symmetry holds because F(s,x)=F(s,x)F(s,-x)=-F(s,x), that is, FF is odd. The vector field FF describes a coupled cell network which has more structure than symmetry alone. For example, the multiplicity 2 eigenvalue λ=4\lambda=4 is not caused by symmetry since all of the irreducible representations of 2×2×2\mathbb{Z}_{2}\times\mathbb{Z}_{2}\times\mathbb{Z}_{2} are one-dimensional [14].

As described in Section 4, we have written a MATLAB® program that computes the bifurcation diagrams of a large class of coupled cell networks systems. We solve Equation (7) on the diamond graph with the results shown in Figure 5. Our program computes one element in each group orbit of branches that is connected to the trivial branch, by recursively following daughter branches of BLIS bifurcations. The bifurcation points are indicated by the circles. If branches cross in this view but there is no circle, then the branches do not cross in ×4\mathbb{R}\times\mathbb{R}^{4} and there is not a bifurcation. We focus on the three highlighted bifurcation points of Figure 5 in the next three examples.

Refer to captionW10W_{10}W1W_{1}W5W_{5}W9W_{9}W6W_{6}W10W_{10}W3W_{3}W4W_{4}W5W_{5}W9W_{9}W0W_{0}W8W_{8}W7W_{7}W5W_{5}W2W_{2}3.173.193.18
Figure 5. The complete bifurcation diagram of Equation (7). The invariant subspace of each branch is indicated. The invariant subspaces are described in Figure 4. This figure shows the bifurcation points that spawn BLIS branches that are neither BSE nor EBL branches, at s=2s=-2 (Example 3.17), s=2.5s=2.5 (Example 3.19), and s=4s=4 (Example 3.18).
Example 3.17.
Refer to captionW6W_{6}W9W_{9}W5W_{5}W1W_{1}W6W_{6}W10W_{10}W5W_{5}
Figure 6. BLIS bifurcations W1W5W_{1}\to W_{5} and W1W6W_{1}\to W_{6} at s=2s=-2, listed as Example 3.17(a) in Figure 1. There are also two pitchfork bifurcations shown that are BSE, EBL, and BLIS: W6W9W_{6}\to W_{9} and W5W10W_{5}\to W_{10}, which are listed as Example 3.17(b) in Figure 1. Only one of the two branches is plotted in each case to avoid clutter. Note that this figure plots a different function of x4x\in\mathbb{R}^{4} vs. ss than the one plotted in Figure 5. Plotting x4x_{4} shows that only one of the W6W_{6} branches has a bifurcation. See Example 3.17.

Continuing the example of the diamond graph, the solution branch in W1W_{1} that bifurcates from the origin at s=0s=0 undergoes two secondary bifurcations, at s=1s=-1 and s=2s=-2, as seen in Figure 5. In this example, we focus on the secondary bifurcation at s=2s=-2, and the tertiary bifurcations of these daughters, as seen in the detailed bifurcation diagram in Figure 6. The bifurcation of the W1W_{1} branch at s=2s=-2 is a BLIS bifurcation that is neither an EBL nor a BSE bifurcation, since W1W_{1} is not a fixed-point subspace, and the critical eigenspace is not one-dimensional. This bifurcation is also described by [5, Theorem 8.2]. It is indicated in Figure 1 as 3.17(a). The daughters are two secondary branches in W5W_{5} and W6W_{6}, and each of these branches undergoes a tertiary bifurcation which is a standard pitchfork bifurcation described by the BSE, EBL and BLIS. These pitchfork bifurcations are indicated in Figure 1 as 3.17(b).

The synchrony basis matrix for W1W_{1} is B1:=[1 1 1 1]TB_{1}:=[1\,1\,1\,1]^{T} with pseudoinverse B1+=14[1 1 1 1]B_{1}^{+}=\frac{1}{4}[1\,1\,1\,1]. Thus B1+LB1=0B_{1}^{+}LB_{1}=0 for the Laplacian matrix (6), and Proposition 3.14 says that F(s,x)=0F(s,x)=0 restricted to W1W_{1} is

FB1(s,y)=sy+y3=0.F_{B_{1}}(s,y)=sy+y^{3}=0.

This has three solutions, y=0y=0 and y=±sy=\pm\sqrt{-s}. The solution y=0y=0 is in W0W_{0}. The latter solutions are in W1W_{1} but not in W0W_{0}. For y=sy=\sqrt{-s}, the mother branch function bm:4b_{m}:\mathbb{R}\to\mathbb{R}^{4} is bm(s)=s(1,1,1,1)b_{m}(s)=\sqrt{-s}(1,1,1,1). We find that D2FB1(s,y)=s+3y2D_{2}F_{B_{1}}(s,y)=s+3y^{2}, and thus D2FB1(s,bm(s))=2sD_{2}F_{B_{1}}(s,b_{m}(s))=-2s. This eigenvalue of this matrix does not change sign, and Proposition 2.8 applies. The eigenvalues of D2F(s,bm(s))D_{2}F(s,b_{m}(s)) are 2s-2s, 1s1-s, and 2s2-s (with multiplicity 2). We focus on the two BLIS bifurcations at s=2s=-2 which have daughter branches in W5W_{5} and W6W_{6}, respectively.

The synchrony basis matrix of W5W_{5} is B5:=[10110100]TB_{5}:=\left[\begin{smallmatrix}1&0&1&1\\ 0&1&0&0\end{smallmatrix}\right]^{T}, and B5+LB5=[1133]B_{5}^{+}LB_{5}=\left[\begin{smallmatrix}1&-1\\ -3&3\end{smallmatrix}\right]. Note that B5+LB5B_{5}^{+}LB_{5} is the Laplacian matrix of the quotient digraph, which has an arrow with weight 1 from cell 2 to cell 1, and an arrow with weight 3 from cell 1 to cell 2. By Proposition 3.14, the system F(s,x)=0F(s,x)=0 restricted to W5W_{5} is

(8) FB5(s,y)=(sy1+y13(y1y2),sy2+y233(y2y1))=0.F_{B_{5}}(s,y)=(sy_{1}+y_{1}^{3}-(y_{1}-y_{2}),sy_{2}+y_{2}^{3}-3(y_{2}-y_{1}))=0.

The Jacobian matrix of FB5F_{B_{5}} is

D2FB5(s,y)=[s+3y12113s+3y223],D_{2}F_{B_{5}}(s,y)=\begin{bmatrix}s+3y_{1}^{2}-1&1\\ 3&s+3y_{2}^{2}-3\end{bmatrix},

and the Jacobian matrix evaluated at the mother branch is

J(s):=D2FB5(s,bm(s))=[2s1132s3].J(s):=D_{2}F_{B_{5}}(s,b_{m}(s))=\begin{bmatrix}-2s-1&1\\ 3&-2s-3\end{bmatrix}.

The eigenvalues of J(s)J(s) are 2s-2s and 2(s+2)-2(s+2). Thus there is a bifurcation at s=2s^{*}=2 and the two components of Theorem 2.12 are

J(2)=[3131],J(2)=[2002].J(-2)=\begin{bmatrix}3&1\\ 3&1\end{bmatrix},\quad J^{\prime}(-2)=\begin{bmatrix}-2&0\\ 0&-2\end{bmatrix}.

We see that N(J(2))=Span({(1,3)})N(J(-2))=\operatorname{Span}(\{(1,-3)\}) is one-dimensional, and the J(2)(1,3)=(2,6)J^{\prime}(-2)(1,-3)=(-2,6) is not in R(J(2))=Span({(1,1)})R(J(-2))=\operatorname{Span}(\{(1,1)\}), so the hypotheses of Theorem 2.12 hold. There is exactly one daughter branch in W5W_{5} bifurcating from the mother branch at s=2s=-2.

Similarly, the synchrony basis matrix of W6W_{6} is B6:=[10100101]TB_{6}:=\left[\begin{smallmatrix}1&0&1&0\\ 0&1&0&1\end{smallmatrix}\right]^{T}. The system F(s,x)=0F(s,x)=0 restricted to W6W_{6} is

(9) FB6(s,y)=(sy1+y132(y1y2),sy2+y232(y2y1))=0.F_{B_{6}}(s,y)=(sy_{1}+y_{1}^{3}-2(y_{1}-y_{2}),sy_{2}+y_{2}^{3}-2(y_{2}-y_{1}))=0.

The Jacobian matrix of FB5F_{B_{5}} is

D2FB6(s,y)=[s+3y12222s+3y222],D_{2}F_{B_{6}}(s,y)=\begin{bmatrix}s+3y_{1}^{2}-2&2\\ 2&s+3y_{2}^{2}-2\end{bmatrix},

and the Jacobian matrix evaluated at the mother branch is

J(s):=D2FB6(s,bm(s))=[2s2222s2].J(s):=D_{2}F_{B_{6}}(s,b_{m}(s))=\begin{bmatrix}-2s-2&2\\ 2&-2s-2\end{bmatrix}.

The eigenvalues of J(s)J(s) are 2s-2s and 2(s+2)-2(s+2). Thus there is a bifurcation at s=2s^{*}=2 and the two components of Theorem 2.12 are

J(2)=[2222],J(2)=[2002].J(-2)=\begin{bmatrix}2&2\\ 2&2\end{bmatrix},\quad J^{\prime}(-2)=\begin{bmatrix}-2&0\\ 0&-2\end{bmatrix}.

We see that N(J(2))=Span({(1,1)})N(J(-2))=\operatorname{Span}(\{(1,-1)\}) is one-dimensional, and the J(2)(1,1)=(2,2)J^{\prime}(-2)(1,-1)=(-2,2) is not in R(J(2))=Span({(1,1)})R(J(-2))=\operatorname{Span}(\{(1,1)\}), so the hypotheses of Theorem 2.12 hold. There is exactly one daughter branch in W6W_{6} bifurcating from the mother branch at s=2s=-2.

Figure 6 shows these two bifurcations. Note that the restricted system (9) has a symmetry (y1,y2)(y2,y1)(y_{1},y_{2})\mapsto(y_{2},y_{1}) and consequently the bifurcation W1W6W_{1}\to W_{6} is a pitchfork bifurcation within the restricted system (9). Within the full system in ×4\mathbb{R}\times\mathbb{R}^{4} there is no such symmetry, so our MATLAB® code follows both branches. The lower W6W_{6} branch in Figure 6 undergoes a bifurcation at s2.4s\approx-2.4, whereas the upper branch has no bifurcation. There is apparently no symmetry of the restricted system (8), and the bifurcation W1W5W_{1}\to W_{5} is transcritical as expected.

Example 3.18.

The trivial branch in System (7) for the diamond graph has a double 0 eigenvalue at s=4s=4. There is not a BSE at (s,x)=(4,0)(s,x)=(4,0) in Figure 5. The irreducible representations of Aut(G)×2\operatorname{Aut}(G)\times\mathbb{Z}_{2} are all one-dimensional, so there is not an EBL bifurcation at that point. However, there are 3 BLIS branches that bifurcate from this point. The critical eigenspace of D2F(4,0)D_{2}F(4,0) is the eigenspace of LL with eigenvalue λ=4\lambda=4, which is E:=Span({(1,1,1,1),(0,1,0,1)})E:=\operatorname{Span}(\{(1,-1,1,-1),(0,1,0,-1)\}). We can use a shortcut to avoid calculations like those in Example 3.17, following Remark 2.13 (1) and (5). For each of the invariant subspaces WiW_{i}, there is a BLIS bifurcation from W0W_{0} to WiW_{i} if the intersection EWiE\cap W_{i} is one-dimensional. Consulting Figure 4, we see that EW2=W2E\cap W_{2}=W_{2}, EW3=W3E\cap W_{3}=W_{3}, and EW5=Span({(1,3,1,1)})E\cap W_{5}=\operatorname{Span}(\{(1,-3,1,1)\}) are all one-dimensional, and EW1=EW4={0}E\cap W_{1}=E\cap W_{4}=\{0\}. It is the case that EW6=W2E\cap W_{6}=W_{2} is one-dimensional, so the BLIS theorem says that there is a daughter branch in W6W_{6}, but this branch is also in W2W_{2}. Thus, we do not need to consider the invariant subspaces that contain W2W_{2}, W3W_{3}, and W5W_{5} and we have found all the BLIS bifurcations: W0W2W_{0}\to W_{2}, W0W3W_{0}\to W_{3}, W0W5W_{0}\to W_{5}. These three bifurcating branches are seen in Figure 5.

Example 3.19.

The BLIS bifurcation of the mother branch W5W_{5} to the daughter branch W8W_{8} at s=2.5s=2.5 is a BSE, but not an EBL bifurcation. For this bifurcation, we will not verify Condition (b) in Theorem 2.12 which is generically true, but focus on Condition (a). The system F(s,x)=0F(s,x)=0 restricted to W5W_{5} was shown in Equation (8). We cannot easily solve that system to find the mother branch function bm(s)b_{m}(s) in closed form, but our MATLAB® program finds numerical solutions. The program also gives a numerical approximation of the bifurcation point, and we can verify by hand that FB5(52,(12,2))=(0,0)F_{B_{5}}(\frac{5}{2},(\frac{1}{\sqrt{2}},-\sqrt{2}))=(0,0). Furthermore, D2FB5(52,(12,2))D_{2}F_{B_{5}}(\frac{5}{2},(\frac{1}{\sqrt{2}},-\sqrt{2})) is nonsingular, so the hypotheses of Proposition 2.8 hold. The system F(s,x)=0F(s,x)=0 restricted to W8W_{8} is FB8(s,y)=0F_{B_{8}}(s,y)=0, where

FB8(s,y)=(sy1+y13(2y1y2y3),sy2+y23(3y22y1y3),sy3+y33(3y32y1y2)).F_{B_{8}}(s,y)=(sy_{1}+y_{1}^{3}-(2y_{1}-y_{2}-y_{3}),sy_{2}+y_{2}^{3}-(3y_{2}-2y_{1}-y_{3}),sy_{3}+y_{3}^{3}-(3y_{3}-2y_{1}-y_{2})).

The Jacobian matrix of FB8F_{B_{8}} is

D2FB8(s,y)=[s+3y122112s+3y223121s+3y323].D_{2}F_{B_{8}}(s,y)=\begin{bmatrix}s+3y_{1}^{2}-2&1&1\\ 2&s+3y_{2}^{2}-3&1\\ 2&1&s+3y_{3}^{2}-3\end{bmatrix}.

While we cannot find the mother branch function bmb_{m}, or J(s)=D2FB8(s,bm(s))J(s)=D_{2}F_{B_{8}}(s,b_{m}(s)), the Jacobian matrix evaluated at the bifurcation point is

J(52)=D2FB8(52,(12,2,12)=[21121121211].\textstyle J(\frac{5}{2})=D_{2}F_{B_{8}}(\frac{5}{2},(\frac{1}{\sqrt{2}},-\sqrt{2},\frac{1}{\sqrt{2}})=\begin{bmatrix}2&1&1\\ 2&\frac{11}{2}&1\\ 2&1&1\end{bmatrix}.

This matrix has a simple zero eigenvalue with an eigenvector of (1,0,2)(1,0,-2). Thus, Condition (a) of Theorem 2.12 is satisfied. While we cannot prove that Condition (b) in the theorem holds, it is true generically, and our MATLAB® program does not check it. The program follows the daughter branches in W8W_{8}, suggesting that Condition (b) is true. This BLIS bifurcation is also a BSE, since the 4×44\times 4 Jacobian matrix evaluated at the bifurcation point has a simple 0 eigenvalue. There is no symmetry, so this is not an EBL bifurcation.

3.3. PDE Examples

The BLIS applies to find primary and secondary bifurcations for many PDE. If all of the hypotheses of the BSE theorem are satisfied except that the critical eigenspace has dimension greater than one, then one can look for one-dimensional invariant subspaces of the critical eigenspace and apply the BLIS there. In this section we provide examples from the semilinear elliptic BVP investigated in [15, 19]. We consider primary bifurcations where the mother branch is the trivial branch defined by bm(s)=0b_{m}(s)=0. In these cases the existence of a daughter branch is proved. For secondary bifurcations, the mother branch function is not known explicitly and the bifurcation is observed numerically rather than proven to exist. The existence of nested invariant subspaces can be proved, and suggests a robust numerical algorithm for following secondary branches.

Consider the PDE

(10) Δu+su+N(u)=0\Delta u+su+N(u)=0

for u:Ωu:\Omega\to\mathbb{R} with 00\,-Dirichlet boundary conditions, where Ω\Omega is a region in n\mathbb{R}^{n}, and N:N:\mathbb{R}\to\mathbb{R} is a nonlinearity which satisfies N(0)=N(0)=0N(0)=N^{\prime}(0)=0. We seek zeros of the function F:×HHF:\mathbb{R}\times H\to H defined by

F(s,u)=u+Δ1(su+N(u)),F(s,u)=u+\Delta^{-1}(su+N(u)),

where HH is the Sobolev space H01,2(Ω)H_{0}^{1,2}(\Omega). We usually choose the subcritical, superlinear nonlinearity N(u)=u3N(u)=u^{3}, as in [15, 19]. In this case and others, regularity theory [8] gives that a zero of FF is twice differentiable and hence a classical solution to the PDE.

The eigenvalue equation associated with PDE (10) is Δu+λu=0\Delta u+\lambda u=0, on the same region with 00\,-Dirichlet boundary conditions. It is well-known [7] that the eigenvalues are real and satisfy 0<λ1<λ2λ30<\lambda_{1}<\lambda_{2}\leq\lambda_{3}\cdots\to\infty . All BSE, EBL, and BLIS bifurcations from the trivial solution are described by Theorem 2.12 with Wm={0}W_{m}=\{0\}. For the current function FF, the operator J(s)=D2F(s,0)J(s)=D_{2}F(s,0) defined in Theorem 2.12 has eigenvalues {1s/λii}\{1-s/\lambda_{i}\mid i\in\mathbb{N}\}. We expect that the branch of trivial solutions has a bifurcation of some sort at (s,u)=(λi,0)(s,u)=(\lambda_{i},0). Suppose we find an invariant subspace WdHW_{d}\subseteq H such that N(J(λi))=Span({x0})N(J(\lambda_{i}))=\operatorname{Span}(\{x_{0}\}), in the notation of Theorem 2.12. We see that all of the conditions of the Theorem hold. The second part of Condition (a) holds because HH is a Hilbert space and Δ1\Delta^{-1} is self-adjoint. Thus, the range of J(λi)J(\lambda_{i}) is the orthogonal complement of the null space, which has codimension one. Condition (b) holds because J(s)=Δ1J^{\prime}(s)=\Delta^{-1}, and therefore J(λi)(x0)=x0/λiR(J(λi))J^{\prime}(\lambda_{i})(x_{0})=x_{0}/\lambda_{i}\not\in R(J(\lambda_{i})).

Example 3.20.

Consider PDE (10). Since λ1\lambda_{1} is simple, the BSE and the BLIS apply at the point (s,u)=(λ1,0)(s,u)=(\lambda_{1},0). There is a unique branch of nontrivial solutions that bifurcates from this point. An explicit example is the PDE

Δu+su+u2=0,\Delta u+su+u^{2}=0,

for which there is a transcritical bifurcation, and for which the EBL does not apply. Note the similarity to Example 3.2.

Example 3.21.

Consider PDE (10) with NN odd. Then F(s,u)=F(s,u)F(s,-u)=-F(s,u), and the bifurcation at (s,u)=(λ1,0)(s,u)=(\lambda_{1},0) is a pitchfork bifurcation. In this case, the BSE, EBL, and BLIS all apply. An explicit example is the PDE

Δu+su+u3=0.\Delta u+su+u^{3}=0.

Note the similarity to Example 3.1.

Example 3.22.

Consider PDE (10) with the square domain, Ω=(0,1)2\Omega=(0,1)^{2}. The eigenvalues and eigenfunctions of the Laplacian can be explicitly computed as

ψn,m(x,y)=sin(nπx)sin(mπy),λn,m=(n2+m2)π2,\psi_{n,m}(x,y)=\sin(n\pi x)\sin(m\pi y),\quad\lambda_{n,m}=(n^{2}+m^{2})\pi^{2},

with n,mn,m\in\mathbb{N}. Note that 0<λ1,1<λ1,2=λ2,10<\lambda_{1,1}<\lambda_{1,2}=\lambda_{2,1}. The EBL applies to the bifurcation at (s,u)=(λ1,2,0)(s,u)=(\lambda_{1,2},0), but BSE does not apply. The group Aut(Ω)\operatorname{Aut}(\Omega) is the symmetry of the square, and that group acts on HH. The fixed point subspaces {uH:u(x,y)=u(y,x)}\{u\in H:u(x,y)=u(y,x)\} and {uH:u(x,y)=u(1x,y)}\{u\in H:u(x,y)=u(1-x,y)\} are FF-invariant and the BLIS applies.

Figure 7. The critical eigenspace VV at the bifurcation point (s,0)(s^{*},0) in Example 3.23, where s=10π2s^{*}=10\pi^{2}. The diagonal lines are intersections of two fixed point subspaces with VV and lead to EBL bifurcations. The horizontal dashed line is the intersection of WdW_{d} with VV and leads to a BLIS bifurcation. The vertical dashed line leads to another similar BLIS bifurcation. The three line types correspond to the three group orbits of bifurcating solutions.
Example 3.23.

Consider bifurcation point (s,u)=(λ1,3,0)(s,u)=(\lambda_{1,3},0) for PDE (10) with the square domain, Ω=(0,1)2\Omega=(0,1)^{2}, with eigenvalues and eigenfunction listed in Example 3.22. The eigenvalue λ1,3=λ3,1\lambda_{1,3}=\lambda_{3,1} is not simple so BSE does not apply. The EBL predicts two group orbits of daughter branches, and the BLIS predicts these two as well as a third group orbit of bifurcating solutions. These are shown in Figure 1 as 3.23(a) and (b), respectively.

The bifurcation of the trivial solution at s=λ1,3s=\lambda_{1,3} has a 2-dimensional critical eigenspace

V=N(D2F(λ1,3,0))=Span({ψ1,3,ψ3,1}).V=N(D_{2}F(\lambda_{1,3},0))=\operatorname{Span}(\{\psi_{1,3},\psi_{3,1}\}).

Figure 7 shows the eigenfunctions in VV that give rise to 8 solution branches, in three group orbits, that bifurcate from the trivial solution at s=λ3,1=10π2s=\lambda_{3,1}=10\pi^{2}. We count this as 8 solution branches following Remark 2.13(7).

The function FF is 𝔻4×2\mathbb{D}_{4}\times\mathbb{Z}_{2}-equivariant, where the symmetry group 𝔻4\mathbb{D}_{4} of the square acts on HH in the natural way and 2\mathbb{Z}_{2} acts as uuu\mapsto-u. The 𝔻4×2\mathbb{D}_{4}\times\mathbb{Z}_{2} action on VV, modulo the kernel of the action, is isomorphic to 𝔻2\mathbb{D}_{2}, and is generated by (ψ1,3,ψ3,1)(ψ3,1,ψ1,3)(\psi_{1,3},\psi_{3,1})\mapsto(\psi_{3,1},\psi_{1,3}) and (ψ1,3,ψ3,1)(ψ1,3,ψ3,1)(\psi_{1,3},\psi_{3,1})\mapsto-(\psi_{1,3},\psi_{3,1}). The EBL says that Span({ψ1,3+ψ3,1})\operatorname{Span}(\{\psi_{1,3}+\psi_{3,1}\}) and Span({ψ1,3ψ3,1})\operatorname{Span}(\{\psi_{1,3}-\psi_{3,1}\}), the diagonals in Figure 7, give rise to branches in two fixed-point subspaces in HH. The bifurcating solutions are either even or odd in each of the reflection lines of the square.

The BLIS shows that there is a bifurcating branch in ×Wd\mathbb{R}\times W_{d}, where

Wd:={uH\displaystyle W_{d}:=\{u\in H\mid\ u(x,1y)=u(x,y) for all (x,y)(0,1)×(0,1) and\displaystyle u(x,1-y)=u(x,y)\text{ for all }(x,y)\in(0,1)\times(0,1)\text{ and }
u(x,23y)=u(x,y) for all (x,y)(0,1)×(0,23)}.\displaystyle u(x,\textstyle\frac{2}{3}-y)=-u(x,y)\text{ for all }(x,y)\in(0,1)\times(0,\textstyle\frac{2}{3})\}.

The intersection of WdW_{d} with VV is Span({ψ1,3})\operatorname{Span}(\{\psi_{1,3}\}), depicted by the horizontal dashed line in Figure 7. The stability of the trivial branch is determined by J(s)=D2Fd(s,0)J(s)=D_{2}F_{d}(s,0), which has a simple eigenvalue equal to 1s/λ1,31-s/\lambda_{1,3} with the critical eigenvector ψ1,3\psi_{1,3}. The symmetry of the square implies that there is another invariant subspace that intersects VV in Span({ψ3,1})\operatorname{Span}(\{\psi_{3,1}\}), leading to another BLIS bifurcation. See [19] for numerical results of these bifurcating branches, as well as other nearby primary and secondary solution branches.

In the next example, the numerically observed bifurcation point is explained by the BLIS, but is not proven to exist.

Example 3.24.

The primary branch predicted by the BLIS in Example 3.23 has a secondary bifurcation, observed numerically and shown in [19, Figure 5]. This bifurcation is not an EBL bifurcation, but it is a BLIS bifurcation as we now explain. The primary branch is in WdW_{d} as implied by the BLIS, but is also observed to be in the invariant fixed point subspace

We,e:={uHu(x,1y)=u(1x,y)=u(x,y) for all (x,y)(0,1)×(0,1)}W_{e,e}:=\{u\in H\mid u(x,1-y)=u(1-x,y)=u(x,y)\text{ for all }(x,y)\in(0,1)\times(0,1)\}

of functions that are even in both the horizontal and vertical directions. The secondary bifurcation shown in [19, Figure 5] is a BLIS bifurcation with nested subspaces We,eWdWe,eW_{e,e}\cap W_{d}\subsetneq W_{e,e}. This is also a BSE bifurcation, since the critical eigenspace is one-dimensional. Schneider has a description of this type of bifurcation in terms of Fourier coefficients of uu in [22, 23].

The invariant subspace WdW_{d} in Example 3.23, based on ψ3,1\psi_{3,1}, is just one of an infinite family of invariant subspaces that are not fixed point subspaces based on ψn,m\psi_{n,m} for all cases where nn or mm is greater than 2. Similar invariant subspaces are even more prominent for the PDE on the cube.

Example 3.25.

In [15] we investigate the PDE Δu+su+u3=0\Delta u+su+u^{3}=0 on the cube (0,π)3(0,\pi)^{3} with 00\,-Dirichlet boundary conditions. Here the eigenvalues of the Laplacian are λ,m,n=2+m2+n2\lambda_{\ell,m,n}=\ell^{2}+m^{2}+n^{2}, with eigenfunctions ψ,m,n(x,y,z)=sin(x)sin(my)sin(nz)\psi_{\ell,m,n}(x,y,z)=\sin(\ell x)\sin(my)\sin(nz). There are 99 group orbits of fixed point subspaces of the symmetry group of the cube acting on the function space HH. These are called symmetry types, denoted S0S_{0} through S98S_{98}. The paper features the bifurcation of the trivial branch at s=λ1,2,3=14s^{*}=\lambda_{1,2,3}=14, for which J(s)J(s^{*}) has a 6-dimensional critical eigenspace.

  1. (a)

    This bifurcation has 4 group orbits of EBL branches, indicated in [15, Figures 23 and 25]. These have symmetry types S11S_{11}, S12S_{12}, S22S_{22}, and S23S_{23}.

  2. (b)

    There are also 2 additional BLIS branches that are not EBL branches. The invariant subspaces are spaces of functions with odd reflection symmetry at z=π/3z=\pi/3 and z=2π/3z=2\pi/3. These are shown in [15, Figure 25] with symmetry type S52S_{52}, with critical eigenfunction ψ1,2,3\psi_{1,2,3}, and the third solution with symmetry type S44S_{44}, with critical eigenfunction ψ1,2,3+ψ2,1,3\psi_{1,2,3}+\psi_{2,1,3}. Thus there are 6 group orbits of BLIS branches that bifurcate from (s,u)=(14,0)(s,u)=(14,0).

  3. (c)

    In addition, there are 13 group orbits of solution branches that are not BLIS branches.

The branches in Parts (a) and (b) are simple to follow, since the critical eigenvector is determined by Theorem 2.12. Following BLIS branches is like shooting fish in a barrel. In contrast, the 13 group orbits of non BLIS branches are found by a random search of the 6-dimensional critical eigenspace. Many hours of computing time were required to find them, and it is possible that some eluded detection. We are reasonably confident we have found all of the branches, with the help of index theory as described in [15].

4. Numerical Implementation

We wrote a MATLAB® program, available at the GitHub repository [18]. It approximates solutions to the system of nn equations

(11) F(s,x)i:=f(s,xi)(Mx)i=0,i{1,2,,n}F(s,x)_{i}:=f(s,x_{i})-(Mx)_{i}=0,\quad i\in\{1,2,\ldots,n\}

for the components of xnx\in\mathbb{R}^{n}. These are equilibrium solutions of Equation (4), with k=1k=1 (so xix_{i}\in\mathbb{R}) and the 1×11\times 1 coupling matrix H=1H=-1. The input data for the program includes the weighted adjacency matrix Mn×nM\in\mathbb{R}^{n\times n}, the nonlinearity f:2f:\mathbb{R}^{2}\rightarrow\mathbb{R}, the permutations that commute with MM, the MM-invariant polydiagonal subspaces, and the containment lattice of the MM-invariant subspaces.

Figures 5 and 6 show bifurcation diagrams with MM chosen to be the Laplacian matrix in Equation (6). We made text files which input the information shown in Figure 4. We chose

(12) f(s,x)=sx+x3,f(s,x)=sx+x^{3},

to yield Equation (7). Other choices of ff, predefined in our MATLAB® code, are

f(s,x)=sx+αx2x3 and f(s,x)=sx+x3βx5,f(s,x)=sx+\alpha x^{2}-x^{3}\text{ and }f(s,x)=sx+x^{3}-\beta x^{5},

where α\alpha and β\beta are parameters which cannot be scaled away. We do not require that ff is odd in the second variable, but the program treats ff odd and ff not odd differently. The MM-invariant synchrony subspaces are FF-invariant for all ff. The MM-invariant anti-synchrony subspaces are FF-invariant if ff is odd (see [21]).

We frequently assume ff satisfies f(s,0)=0f(s,0)=0, D1f(s,x)=xD_{1}f(s,x)=x, and D2f(s,0)=sD_{2}f(s,0)=s for all ss, as in the above examples. This has the advantage that the stability of the trivial branch is easy to compute. For the trivial branch, with bm(s)=0b_{m}(s)=0 and Wd=nW_{d}=\mathbb{R}^{n}, the stability matrix J(s)=sInMJ(s)=sI_{n}-M is easily analyzed based on the eigenvalues of MM (see Equation (2)). Our program works for any smooth ff, although if f(s,0)0f(s,0)\neq 0, then an approximate solution must be supplied to the program as a starting point for the first branch. For example we have tested

f(s,x)=(s21)x±x3,f(s,x)=sx+x3+1.f(s,x)=(s^{2}-1)x\pm x^{3},\quad f(s,x)=sx+x^{3}+1.

We now describe the general algorithms that, when implemented, can generate numerical bifurcation results for networks such as those displayed in Figures 5 and 6. There are four essential processes involved:

  1. (1)

    Compute the lattice of invariant subspaces.

  2. (2)

    Follow branches, given a starting point, the invariant subspace, and initial direction.

  3. (3)

    Detect bifurcation points, which are the starting points of daughter branches.

  4. (4)

    Use BLIS to find the invariant subspace and initial direction of each daughter branch.

For the first process, we start with the matrix MM in (11) (typically the graph Laplacian). The algorithm described in [21] and implemented on that article’s GitHub repository [20] finds all of the MM-invariant polydiagonal subspaces and the lattice showing which subspaces are contained in which others. For example, the information in Figure 4 can be computed with this program.

The weighted analytic matrix MM, invariant subspaces, lattice of inclusion, and Aut(G)\operatorname{Aut}(G) information is passed to the MATLAB® program by four text files, as described in the user’s manual available at [18].

Given these four text files as inputs, the MATLAB® code posted on the current paper’s GitHub repository [18] computes the bifurcation diagrams showing all of the solutions to Equation (11) that are within the strip sminssmaxs_{\text{min}}\leq s\leq s_{\text{max}} and connected by a series of BLIS bifurcations to the starting solution, usually x=0x=0, s=smins=s_{\text{min}}. The input parameters are changed by editing the MATLAB® code.

Branch following is achieved by employing a new version of the so-called tGNGA (tangent gradient Newton Galerkin Algorithm) described in [15]. The added feature is that the computations are done within the invariant subspace. That is, we solve Equation (5) (with H=1H=-1) rather than Equation (11). This speeds up the computations greatly, especially in examples where nn is large and dd is small.

The input to the tGNGA is the invariant subspace WmW_{m} of the branch, a single point (s~,x~)×Wm(\tilde{s},\tilde{x})\in\mathbb{R}\times W_{m}, and a tangent vector to the branch in (s~,x~)×Wm(\tilde{s},\tilde{x})\in\mathbb{R}\times W_{m}. Typically, the first branch has x~=0\tilde{x}=0 and Wm={0}W_{m}=\{0\}. After that, the initial point is a bifurcation point on a previously computed branch.

The branch following algorithm needs to stop the branch at a bifurcation point when it is the daughter branch. If the branch continues past the bifurcation point, then it can potentially continue past another bifurcation point and cause an infinite loop in the MATLAB® program.

Detecting bifurcation points on a mother branch in ×Wm\mathbb{R}\times W_{m} is done using the BLIS. The Morse Index, which is a non-negative integer, is an important feature in our GNGA [14, 15, 19]. It is replaced by a vector of non-negative integers in our new code. Let 𝒟\mathcal{D} be the set of FF-invariant subspaces of System (11) that contain WmW_{m}. As described in [21], 𝒟\mathcal{D} is different depending on whether or not ff is odd. These are the possible daughter subspaces. At each computed point on the mother branch, the eigenvalues of the Jacobian, restricted to each W𝒟W\in\mathcal{D}, are computed, and a list of the number of eigenvalues with negative real part is recorded. This is the signature list. Recall that the matrix MM in Equation (11) need not be symmetric, so the eigenvalues of the Jacobian are not necessarily real. If any of the numbers on the signature list changes between two points, the secant method is used to find the point where an eigenvalue has zero real part. At a fold point, the signature of the Jacobian, restricted to WmW_{m}, changes by 1 and this is not a bifurcation point. If any other number on the signature list changes by 1, the point where the corresponding eigenvalue is 0 is a BLIS bifurcation point, and the corresponding critical eigenvector is computed.

The invariant subspace of the daughter branch is the smallest W𝒟W\in\mathcal{D} that contains the critical eigenvector. Care must be taken to follow only one daughter branch if several of them are related by the Aut(G)×2\operatorname{Aut}(G)\times\mathbb{Z}_{2} symmetry. Then the bifurcation point, the invariant subspace, and the critical eigenvector are put into a queue to start the daughter branch with the tGNGA. We do not check Condition (b) in Theorem 2.12. This condition is generically true, and it is observed to be true in all of the bifurcations we have computed. In this way, we follow one branch in each group orbit of branches that are connected to the first branch (usually the trivial branch) by a sequence of BLIS bifurcations.

5. Conclusion

We have presented the Bifurcation Lemma for Invariant Subspaces (BLIS), which describes most generic bifurcations of solutions to F(s,x)=0F(s,x)=0, for F:×XXF:\mathbb{R}\times X\to X. The space XX can be the Euclidean space n\mathbb{R}^{n}, or more generally a Banach space. Thus, the theory applies to PDE. The theory of Bifurcation from Simple Eigenvalues (BSE) describes the simplest case. The BLIS is BSE applied to nested invariant subspaces of FF. The BLIS allows the analysis of some bifurcations with multiple 0 eigenvalues. The Equivariant Branching Lemma (EBL) has been a powerful tool for describing bifurcations where high multiplicity of a zero eigenvalue is caused by the symmetry of the system. The EBL is a special case of the BLIS. Many coupled cell networks have invariant subspaces beyond those caused by symmetry, and the BLIS can describe bifurcations in such systems. We have written a freely available MATLAB® program that computes the solution branches in a class of coupled cell networks.

While the BLIS does not describe all steady state bifurcations, in many examples most or all of the bifurcating branches are predicted by this simple theorem which is easy to apply.

References

  • [1] Manuela Aguiar and Ana Dias. Synchrony and antisynchrony in weighted networks. SIAM Journal on Applied Dynamical Systems, 20(3):1382–1420, 2021.
  • [2] Manuela A. D. Aguiar and Ana Paula S. Dias. The lattice of synchrony subspaces of a coupled cell network: characterization and computation algorithm. J. Nonlinear Sci., 24(6):949–996, 2014.
  • [3] G. Cicogna. Symmetry breakdown from bifurcation. Lett. Nuovo Cimento (2), 31(17):600–602, 1981.
  • [4] Michael G. Crandall and Paul H. Rabinowitz. Bifurcation from simple eigenvalues. J. Functional Analysis, 8:321–340, 1971.
  • [5] Alessio Franci, Martin Golubitsky, Ian Stewart, Anastasia Bizyaeva, and Naomi E. Leonard. Breaking indecision in multiagent, multioption dynamics. SIAM Journal on Applied Dynamical Systems, 22(3):1780–1817, 2023.
  • [6] Karin Gatermann. Computation of bifurcation graphs. In Exploiting symmetry in applied and numerical analysis (Fort Collins, CO, 1992), volume 29 of Lectures in Appl. Math., pages 187–201. Amer. Math. Soc., Providence, RI, 1993.
  • [7] D. Gilbarg and N.S. Trudinger. Elliptic Partial Differential Equations of Second Order. Classics in Mathematics. Springer Berlin Heidelberg, 2001.
  • [8] David Gilbarg and Neil S. Trudinger. Generalized Solutions and Regularity, pages 177–218. Springer Berlin Heidelberg, Berlin, Heidelberg, 2001.
  • [9] Martin Golubitsky and Reiner Lauterbach. Bifurcations from synchrony in homogeneous networks: Linear theory. SIAM Journal on Applied Dynamical Systems, 8(1):40–75, 2009.
  • [10] Martin Golubitsky and David G. Schaeffer. Singularities and groups in bifurcation theory. Vol. I, volume 51 of Applied Mathematical Sciences. Springer-Verlag, New York, 1985.
  • [11] Martin Golubitsky and Ian Stewart. Dynamics and Bifurcation in Networks: Theory and Applications of Coupled Differential Equations. Society for Industrial and Applied Mathematics, Philadelphia, PA, 2023.
  • [12] Martin Golubitsky, Ian Stewart, and David G. Schaeffer. Singularities and groups in bifurcation theory. Vol. II, volume 69 of Applied Mathematical Sciences. Springer-Verlag, New York, 1988.
  • [13] Hiroko Kamei and Peter J. A. Cock. Computation of balanced equivalence relations and their lattice for a coupled cell network. SIAM J. Appl. Dyn. Syst., 12(1):352–382, 2013.
  • [14] John M. Neuberger, Nándor Sieben, and James W. Swift. Automated bifurcation analysis for nonlinear elliptic partial difference equations on graphs. Internat. J. Bifur. Chaos Appl. Sci. Engrg., 19(8):2531–2556, 2009.
  • [15] John M. Neuberger, Nándor Sieben, and James W. Swift. Newton’s method and symmetry for semilinear elliptic PDE on the cube. SIAM J. Appl. Dyn. Syst., 12(3):1237–1279, 2013.
  • [16] John M. Neuberger, Nándor Sieben, and James W. Swift. Invariant Synchrony Subspaces of Sets of Matrices (companion web site), 2019. https://jan.ucc.nau.edu/ns46/invariant.
  • [17] John M. Neuberger, Nándor Sieben, and James W. Swift. Invariant synchrony subspaces of sets of matrices. SIAM J. Appl. Dyn. Syst., 19(2):964–993, 2020.
  • [18] John M. Neuberger, Nándor Sieben, and James W. Swift. Github repository, https://github.com/jwswift/BLIS/, 2023.
  • [19] John M. Neuberger and James W. Swift. Newton’s method and Morse index for semilinear elliptic PDEs. Internat. J. Bifur. Chaos Appl. Sci. Engrg., 11(3):801–820, 2001.
  • [20] Eddie Nijholt, Nándor Sieben, and James W. Swift. Github repository, https://github.com/jwswift/Anti-Synchrony-Subspaces/, 2022.
  • [21] Eddie Nijholt, Nándor Sieben, and James W. Swift. Invariant synchrony and anti-synchrony subspaces of weighted networks. J. Nonlinear Sci., 33(4):38, 2023. Id/No 63.
  • [22] I. Schneider and J.-Y. Dai. Symmetry groupoids for pattern-selective feedback stabilization of the Chafee-Infante equation. Chaos, 33(7):Paper No. 073141, 9, 2023.
  • [23] Isabelle Anne Nicole Schneider. Symmetry Groupoids in Dynamical Systems. Habilitation, 2022.
  • [24] Pedro Soares. Synchrony branching lemma for regular networks. SIAM J. Appl. Dyn. Syst., 16(4):1869–1892, 2017.
  • [25] A. Vanderbauwhede. Local bifurcation and symmetry, volume 75 of Research Notes in Mathematics. Pitman (Advanced Publishing Program), Boston, MA, 1982.