A Bifurcation Lemma for Invariant Subspaces
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 subspaces2020 Mathematics Subject Classification:
34A34, 34C23, 35J61, 37C79, 37C811. Introduction
We present a Bifurcation Lemma for Invariant Subspaces (BLIS). The BLIS proves the existence of bifurcating solution branches of a nonlinear equation with a single parameter . 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.
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 , where is an open interval and is a Banach space. A subspace of is -invariant if for all .
This extends the standard definition that is an -invariant subspace for provided .
Example 2.2.
The trivial subspace is -invariant if and only if for all . Also, the full subspace is -invariant for any .
We often get -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 acts linearly on a Banach space by the function . We say that is -equivariant if for all , . The stabilizer of , or the isotropy subgroup of with respect to , is . An isotropy subgroup of is for some . A subgroup of is not necessarily an isotropy subgroup of the group action. For every subgroup of , the fixed point subspace of is
The point stabilizer of a set is
The group action on extends to an action on subsets of . If is an -invariant subspace, then is also an -invariant subspace for all . The -orbit of an -invariant subspace is .
Proposition 2.3.
Given a subgroup of a compact Lie group that acts linearly on a Banach space, is an -invariant subspace if is -equivariant.
Proof.
First of all, is a subspace of because acts linearly. Assume . Then for all . Thus . ∎
Example 2.4.
Let be defined by , where is the space of infinitely differentiable real-valued functions on . The group acts on by . Note that is -equivariant because . The -invariant fixed point subspace is the subspace of even functions.
Recall that a simple -curve is the image of an injective -function with non-vanishing derivative. Note that a simple -curve does not cross itself, and it has no cusps or corners.
Definition 2.5.
Let be as in Definition 2.1. A solution branch of , or simply a branch, is a simple -curve that is a subset of . A -branch is a solution branch contained in , where is an -invariant subspace of .
Example 2.6.
The function defined by has two solution branches, with equations and . The L-shaped curve is not a solution branch because it has a corner at .
Definition 2.7.
A point in a solution branch of is a bifurcation point if every neighborhood of in contains a zero of that is not in the solution branch.
Typically, a bifurcation point is the intersection of two or more branches. In the most recent example, 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 to 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 be an open interval, be a Banach space, and be a -function. Suppose is an -invariant subspace of , satisfies , and is nonsingular. Then there exists a neighborhood of in , and a -function defined on an open interval containing such that and
where
(1) |
We call the mother branch function and its graph the mother branch.
Proof.
This is the Implicit Function Theorem applied to the restriction of to . The only subtlety is that is a neighborhood of in . ∎
Remark 2.9.
If is an invariant subspace, then is vacuously nonsingular. Here the mother branch function is for all , and the mother branch is called the trivial branch.
The BLIS describes a bifurcation of the mother branch of Proposition 2.8 at to a daughter branch. This is a bifurcation from a simple eigenvalue for the restricted function , where 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 , while [4] assumes . In our theorem, the bifurcation is at , whereas [4] assumes the bifurcation point is .
Theorem 2.10 (BSE, Theorem 1.7 of [4]).
Let be a Banach space and an open interval containing . Let be a -function with these properties:
(a) for .
(b) The partial derivatives , , , and exist and are continuous.
(c) is one-dimensional, and has codimension 1.
(d) , where
If is any complement of in , then there is a neighborhood of in , an interval , and -functions , such that , and
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 to . Choose such that . We apply [4, Theorem 1.7] to defined by . Hypothesis (b) in [4, Theorem 1.7] does not require to be continuous, and the conclusion is that and are continuous. In a remark after the theorem they state that if is continuous, then and are , 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 , with and for , and .
Theorem 2.12 (Bifurcation Lemma for Invariant Subspaces).
Assume that the hypotheses of Proposition 2.8 hold. Let and be the mother branch function and mother branch. Assume be nested -invariant subspaces. Let be the restriction of . Assume , satisfies the following conditions:
(a) is one-dimensional, and has codimension one.
(b) , where .
If is any complement of in , then there is a neighborhood of in and -functions and with and such that
where
is the so-called daughter branch. Furthermore, . That is, there are exactly two branches in that contain ; the mother branch and the daughter branch .
Proof.
Since is -invariant, the restriction is well-defined. We will show that the hypotheses of Theorem 2.10 apply to the function defined by
Condition (a) holds, since , by the definition of the mother branch function . The restriction to is allowed since . Condition (b) holds: is since is and is . Conditions (c) and (d) require the computation of and . The Jacobian is
and the stability of the solution at any on the mother branch is determined by
(2) |
While 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
and is equivalent to Condition (c) of Theorem 2.10. Condition (b) of Theorem 2.12 concerns
and is equivalent to Condition (d) of Theorem 2.10.
Thus, Theorem 2.10 holds for the function , and for the functions and defined in that theorem,
for . By the definition of in terms of ,
for , 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 . Finally, the point on the daughter branch when is . There is no other point on the daughter branch in , since , and , and is in . Thus, the only point of intersection of and is . ∎
Remark 2.13.
We give several observations.
-
(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 crosses 0 with nonzero speed at . This leads to an informal statement of the BLIS:
Suppose are invariant subspaces of , with for . Let . If
then generically has exactly two solution branches in that contain ; the mother branch which is in , 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)
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 acts on a Banach space , that is -equivariant, and that . Let satisfy and let . If
then generically a branch of solutions in bifurcates from the mother branch, which is in , at .
-
(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 be the finite dimensional phase space of a network with the admissible map , defined in [5]. Let be the space of fully synchronous states, and let be the synchrony subspace for some balanced coloring . Let , , and let . If
then generically a unique branch of equilibria with synchrony pattern bifurcates from .
-
(4)
There might be many other branches bifurcating from . When we restrict to branches in , however, there are exactly two branches (the mother and daughter branches).
- (5)
-
(6)
If is finite dimensional, then the two parts of Condition (a) of Theorem 2.12 are equivalent. That is, we only need to check that is one-dimensional.
-
(7)
Rather than having one daughter branch that crosses the mother branch at the bifurcation point, we often consider two daughter branches and , born at the bifurcation, defined by
Note that for the parametric curve while for . Neither of these branches include , and
The MATLAB® code, described in Section 4 and freely available at [18], follows the branches and separately, starting from their birth at . At a pitchfork bifurcation where the two daughter branches are conjugate by a symmetry of the system, the code only follows .
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 . We will write examples of ODES, , and look for the equilibrium solutions which satisfy , for functions . In such cases, in the statement of Theorem 2.12 is a matrix.
Example 3.1.
The ODE
has a pitchfork bifurcation at . In this case, . Factoring shows that the mother branch has the equation , and the daughter branch has the equation .
The mother branch function is for all , and . The bifurcation point is , and , and . The system has the symmetry , so both and are fixed point subspaces. The bifurcation is BSE, EBL, and BLIS. See Figure 2(a).
Example 3.2.
The ODE
has a so-called transcritical bifurcation at . In this case, . Factoring shows that the mother branch has the equation , and the daughter branch has the equation .
The mother branch function is for all , and . The bifurcation point is , and , and . The trivial subspace is -invariant since for all , The bifurcation is BSE and BLIS. See Figure 2(b).
Example 3.3.
The ODE
has a different kind of transcritical bifurcation at . In this case, the only -invariant subspace is , so this is not a BLIS bifurcation. The function has no symmetry, so this is not an EBL bifurcation. Factoring shows that there are two branches, and . Arbitrarily choose one of these branches, and define . The BSE Theorem 2.10 and BLIS Theorem 2.12 both apply to with . The relevant calculations are that , so and . The critical eigenvector is , so . It is clear from factoring that the two branches of are and , so
Theorem 1 of [4] describes how the original function , which a branch of nontrivial solutions, has a bifurcation. Thus we say that has a BSE bifurcation, as indicated in Figure 1.
Although Theorem 2.12 applies to the reduced function , it does not apply directly to . Since the original function 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 is added, then the system becomes , which has a qualitatively different bifurcation diagram of vs. for fixed , , or . In contrast, the transcritical bifurcation in Example 3.2 is generic for the class of systems with , which is a natural constraint for some models. See Figure 2(c).
Example 3.4.
Example 3.5.
Consider the ODE
The mother branch function is , and the stability of the trivial branch is determined by . Condition (a) of Theorem 2.12 holds, with , , , and the critical eigenvector . However the nondegeneracy Condition (b) of Theorem 2.12 does not hold, since and thus . Similarly, the BSE does not hold for this example, since Condition (d) of Theorem 2.10 is false. The function 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 that contain . Similar examples of for which Condition (a) of the BLIS holds, but Condition (b) does not hold, can be manufactured for any . See Figure 2(e).
3.2. Network Examples
We apply the BLIS to systems of networks of coupled cells defined by for . The ODE for each cell with phase space is
(3) |
where describes the internal dynamics of each identical cell, is the coupling matrix, and is the weighted adjacency matrix of the network. Note that we multiply a real matrix with a matrix whose entries come from , resulting in a matrix , which we identify with .
In order to understand the BLIS bifurcations, we first describe the subspaces that are invariant under System (3) for a fixed with any and . These invariant subspaces are related to the -invariant polydiagonal subspaces of , as described in [21]. Here we show that the dynamics on one of these invariant subspaces is also described by System (3) but with coupled cells, where the matrix is replaced by a smaller matrix. We can think of any matrix as the adjacency matrix of a weighted digraph with vertices, and then the new matrix is the weighted adjacency matrix of the weighted quotient digraph with 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 , a polydiagonal basis matrix is a matrix that satisfies the following three conditions:
-
(1)
;
-
(2)
every row of has at most one non-zero entry;
-
(3)
is in reduced row-echelon form.
If, in addition, every row of has an entry of 1, then is called a synchrony basis matrix. Otherwise, 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 .
Proposition 3.7.
A nontrivial polydiagonal subspace of is the column space of a unique polydiagonal matrix. Furthermore, is a synchrony/anti-synchrony subspace if and only if 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 can both be defined by equations of the form or (including when ). Any set of equations defining an anti-synchrony subspace has at least one equation of the form . Conditions (1) and (3) in Definition 3.6 ensure that implies .
Note that the trivial subspace 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 is the synchrony basis matrix for . The corresponding tagged partition, as defined in [21], is with the empty partial involution.
The matrix is the anti-synchrony basis matrix for . The corresponding tagged partition is with and . Note that is not in the domain of the partial involution.
Let be a polydiagonal basis matrix. Then is -invariant if and only if . See [17] for an efficient way to find the -invariant synchrony subspaces of . Note that the columns of are orthogonal due to Condition (2), and is a diagonal matrix, with equal to the number of nonzero components in the th column of . Condition (1) implies that . The Moore-Penrose inverse (pseudoinverse)
of is easily computed since is diagonal and nonsingular. The pseudoinverse satisfies and is the projection of onto .
Example 3.10.
Let . Then , so
is the pseudoinverse of .
We now consider the dynamically invariant subspaces of System (3), which are precisely the -invariant subspaces. Proposition 4.1 of [21] describes these subspaces. If and are fixed, then the nontrivial subspaces of that are -invariant for all odd are precisely the subspaces of the form
for some polydiagonal basis matrix such that is -invariant. The trivial subspace -invariant whenever . Furthermore, if is not odd and is a synchrony basis matrix, then is -invariant.
Note that is a product similar to in System (3). If , then . If , then , which is the tensor product of with the polydiagonal subspace [21].
Example 3.11.
Let and , so that . We find that
so . A typical element of is , which we can abbreviate to , as in Figure 3.
We now describe how to restrict the System (3) to an invariant subspace . To simplify the exposition we first consider a network of uncoupled cells with no parameter.
Proposition 3.12.
Let be a polydiagonal basis matrix, , and defined by . If is odd or is a synchrony basis matrix, then defined by satisfies
Proof.
Let
be the set of indices such that the th row of has a nonzero element. Define and such that
If , then . If , then and .
If is odd, then for and for . If is a synchrony basis matrix, then for all . In either case,
Let be the number of nonzero elements in the th column of . Note that is the size of the set . The pseudoinverse of has components
Then
since the sum has nonzero terms, all equal to . ∎
Remark 3.13.
Proposition 3.12 is the motivation for the definition of a polydiagonal basis matrix. Consider , which is not a polydiagonal basis matrix since even though is a polydiagonal subspace which is -invariant for the uncoupled system in Proposition 3.12 if . Since and in general , the single reduced cell may not have the internal dynamics .
We now use Proposition 3.12 to write in terms of , including the parameter .
Proposition 3.14.
Let be a polydiagonal basis matrix, be an -invariant polydiagonal subspace, , , and be defined by
(4) |
Assume that is odd or is a synchrony basis matrix. The function defined by satisfies
(5) |
Proof.
The internal dynamics is handled in Proposition 3.12, and the inclusion of the parameter in both and has no effect on the result. Recall that means that . Thus,
∎
Note the similarity of Equations (4) and (5). The internal dynamics is the same in both equations, but in Equation (4) is replaced by in Equation (5). We interpret as the adjacency matrix of a weighted digraph with vertices, and is the adjacency matrix of the weighted quotient digraph with 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 component equations of are restricted to a polydiagonal subspace . It is clear that only equations are essential and the rest hold automatically, as demonstrated in the next example.
Example 3.15.
Let
Note that , so is -invariant. System (3) is , where
Assume is odd. It follows that is an -invariant subspace, and the ODE can be restricted to by setting , and . We find that and , so only the first and second components are needed. Proposition 3.14 formalizes this procedure. The pseudoinverse of is computed in Example 3.10, which yields the adjacency matrix
of the weighted quotient digraph. System (3) restricted to is , where
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.
![]() |
![]() |
![]() |
(a) | (b) | (c) |
We have investigated System (3) for all of the connected graphs with 4 or fewer vertices, with 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 be the 4-vertex diamond graph, shown in Figure 4. The graph Laplacian matrix of is
(6) |
The lattice of -invariant subspaces are shown in Figure 4.


Assume that the state space of each cell has dimension , and the internal dynamics is . We set , and . Note that we identify the matrix with a real number. Thus, System (3) becomes
(7) |
Thus we have a function whose component functions are Equation (7). We search for solutions to , and we want to understand the bifurcations. As proved in [21], the -invariant subspaces are the -invariant subspaces. These are shown in Figure 4 .
The eigenvalues of are with eigenvector , with eigenvector , and with an eigenspace spanned by and .
The vector field is -equivariant, where is the symmetry group of the graph . The extra symmetry holds because , that is, is odd. The vector field describes a coupled cell network which has more structure than symmetry alone. For example, the multiplicity 2 eigenvalue is not caused by symmetry since all of the irreducible representations of 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 and there is not a bifurcation. We focus on the three highlighted bifurcation points of Figure 5 in the next three examples.
Example 3.17.
Continuing the example of the diamond graph, the solution branch in that bifurcates from the origin at undergoes two secondary bifurcations, at and , as seen in Figure 5. In this example, we focus on the secondary bifurcation at , and the tertiary bifurcations of these daughters, as seen in the detailed bifurcation diagram in Figure 6. The bifurcation of the branch at is a BLIS bifurcation that is neither an EBL nor a BSE bifurcation, since 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 and , 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 is with pseudoinverse . Thus for the Laplacian matrix (6), and Proposition 3.14 says that restricted to is
This has three solutions, and . The solution is in . The latter solutions are in but not in . For , the mother branch function is . We find that , and thus . This eigenvalue of this matrix does not change sign, and Proposition 2.8 applies. The eigenvalues of are , , and (with multiplicity 2). We focus on the two BLIS bifurcations at which have daughter branches in and , respectively.
The synchrony basis matrix of is , and . Note that 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 restricted to is
(8) |
The Jacobian matrix of is
and the Jacobian matrix evaluated at the mother branch is
The eigenvalues of are and . Thus there is a bifurcation at and the two components of Theorem 2.12 are
We see that is one-dimensional, and the is not in , so the hypotheses of Theorem 2.12 hold. There is exactly one daughter branch in bifurcating from the mother branch at .
Similarly, the synchrony basis matrix of is . The system restricted to is
(9) |
The Jacobian matrix of is
and the Jacobian matrix evaluated at the mother branch is
The eigenvalues of are and . Thus there is a bifurcation at and the two components of Theorem 2.12 are
We see that is one-dimensional, and the is not in , so the hypotheses of Theorem 2.12 hold. There is exactly one daughter branch in bifurcating from the mother branch at .
Figure 6 shows these two bifurcations. Note that the restricted system (9) has a symmetry and consequently the bifurcation is a pitchfork bifurcation within the restricted system (9). Within the full system in there is no such symmetry, so our MATLAB® code follows both branches. The lower branch in Figure 6 undergoes a bifurcation at , whereas the upper branch has no bifurcation. There is apparently no symmetry of the restricted system (8), and the bifurcation is transcritical as expected.
Example 3.18.
The trivial branch in System (7) for the diamond graph has a double 0 eigenvalue at . There is not a BSE at in Figure 5. The irreducible representations of 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 is the eigenspace of with eigenvalue , which is . 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 , there is a BLIS bifurcation from to if the intersection is one-dimensional. Consulting Figure 4, we see that , , and are all one-dimensional, and . It is the case that is one-dimensional, so the BLIS theorem says that there is a daughter branch in , but this branch is also in . Thus, we do not need to consider the invariant subspaces that contain , , and and we have found all the BLIS bifurcations: , , . These three bifurcating branches are seen in Figure 5.
Example 3.19.
The BLIS bifurcation of the mother branch to the daughter branch at 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 restricted to was shown in Equation (8). We cannot easily solve that system to find the mother branch function 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 . Furthermore, is nonsingular, so the hypotheses of Proposition 2.8 hold. The system restricted to is , where
The Jacobian matrix of is
While we cannot find the mother branch function , or , the Jacobian matrix evaluated at the bifurcation point is
This matrix has a simple zero eigenvalue with an eigenvector of . 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 , suggesting that Condition (b) is true. This BLIS bifurcation is also a BSE, since the 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 . 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) |
for with -Dirichlet boundary conditions, where is a region in , and is a nonlinearity which satisfies . We seek zeros of the function defined by
where is the Sobolev space . We usually choose the subcritical, superlinear nonlinearity , as in [15, 19]. In this case and others, regularity theory [8] gives that a zero of is twice differentiable and hence a classical solution to the PDE.
The eigenvalue equation associated with PDE (10) is , on the same region with -Dirichlet boundary conditions. It is well-known [7] that the eigenvalues are real and satisfy . All BSE, EBL, and BLIS bifurcations from the trivial solution are described by Theorem 2.12 with . For the current function , the operator defined in Theorem 2.12 has eigenvalues . We expect that the branch of trivial solutions has a bifurcation of some sort at . Suppose we find an invariant subspace such that , 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 is a Hilbert space and is self-adjoint. Thus, the range of is the orthogonal complement of the null space, which has codimension one. Condition (b) holds because , and therefore .
Example 3.20.
Consider PDE (10). Since is simple, the BSE and the BLIS apply at the point . There is a unique branch of nontrivial solutions that bifurcates from this point. An explicit example is the PDE
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.
Example 3.22.
Consider PDE (10) with the square domain, . The eigenvalues and eigenfunctions of the Laplacian can be explicitly computed as
with . Note that . The EBL applies to the bifurcation at , but BSE does not apply. The group is the symmetry of the square, and that group acts on . The fixed point subspaces and are -invariant and the BLIS applies.
Example 3.23.
Consider bifurcation point for PDE (10) with the square domain, , with eigenvalues and eigenfunction listed in Example 3.22. The eigenvalue 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 has a 2-dimensional critical eigenspace
Figure 7 shows the eigenfunctions in that give rise to 8 solution branches, in three group orbits, that bifurcate from the trivial solution at . We count this as 8 solution branches following Remark 2.13(7).
The function is -equivariant, where the symmetry group of the square acts on in the natural way and acts as . The action on , modulo the kernel of the action, is isomorphic to , and is generated by and . The EBL says that and , the diagonals in Figure 7, give rise to branches in two fixed-point subspaces in . 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 , where
The intersection of with is , depicted by the horizontal dashed line in Figure 7. The stability of the trivial branch is determined by , which has a simple eigenvalue equal to with the critical eigenvector . The symmetry of the square implies that there is another invariant subspace that intersects in , 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 as implied by the BLIS, but is also observed to be in the invariant fixed point subspace
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 . 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 in [22, 23].
The invariant subspace in Example 3.23, based on , is just one of an infinite family of invariant subspaces that are not fixed point subspaces based on for all cases where or 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 on the cube with -Dirichlet boundary conditions. Here the eigenvalues of the Laplacian are , with eigenfunctions . There are 99 group orbits of fixed point subspaces of the symmetry group of the cube acting on the function space . These are called symmetry types, denoted through . The paper features the bifurcation of the trivial branch at , for which has a 6-dimensional critical eigenspace.
-
(a)
This bifurcation has 4 group orbits of EBL branches, indicated in [15, Figures 23 and 25]. These have symmetry types , , , and .
-
(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 and . These are shown in [15, Figure 25] with symmetry type , with critical eigenfunction , and the third solution with symmetry type , with critical eigenfunction . Thus there are 6 group orbits of BLIS branches that bifurcate from .
-
(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 equations
(11) |
for the components of . These are equilibrium solutions of Equation (4), with (so ) and the coupling matrix . The input data for the program includes the weighted adjacency matrix , the nonlinearity , the permutations that commute with , the -invariant polydiagonal subspaces, and the containment lattice of the -invariant subspaces.
Figures 5 and 6 show bifurcation diagrams with chosen to be the Laplacian matrix in Equation (6). We made text files which input the information shown in Figure 4. We chose
(12) |
to yield Equation (7). Other choices of , predefined in our MATLAB® code, are
where and are parameters which cannot be scaled away. We do not require that is odd in the second variable, but the program treats odd and not odd differently. The -invariant synchrony subspaces are -invariant for all . The -invariant anti-synchrony subspaces are -invariant if is odd (see [21]).
We frequently assume satisfies , , and for all , 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 and , the stability matrix is easily analyzed based on the eigenvalues of (see Equation (2)). Our program works for any smooth , although if , then an approximate solution must be supplied to the program as a starting point for the first branch. For example we have tested
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)
Compute the lattice of invariant subspaces.
-
(2)
Follow branches, given a starting point, the invariant subspace, and initial direction.
-
(3)
Detect bifurcation points, which are the starting points of daughter branches.
-
(4)
Use BLIS to find the invariant subspace and initial direction of each daughter branch.
For the first process, we start with the matrix in (11) (typically the graph Laplacian). The algorithm described in [21] and implemented on that article’s GitHub repository [20] finds all of the -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 , invariant subspaces, lattice of inclusion, and 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 and connected by a series of BLIS bifurcations to the starting solution, usually , . 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 ) rather than Equation (11). This speeds up the computations greatly, especially in examples where is large and is small.
The input to the tGNGA is the invariant subspace of the branch, a single point , and a tangent vector to the branch in . Typically, the first branch has and . 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 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 be the set of -invariant subspaces of System (11) that contain . As described in [21], is different depending on whether or not is odd. These are the possible daughter subspaces. At each computed point on the mother branch, the eigenvalues of the Jacobian, restricted to each , 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 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 , 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 that contains the critical eigenvector. Care must be taken to follow only one daughter branch if several of them are related by the 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 , for . The space can be the Euclidean space , 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 . 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.