1. Introduction and main results
The setting of this paper is the systems of reaction-diffusion equations of the form
|
|
|
(1.1) |
where are space and time, respectively, and are the gradients of the function , and has the form , where is the set of all matrices. Such a system is of the activator-inhibitor type and will be referred to as a skew-gradient system [18].
A traveling wave solution of one variable is a solution to (1.1) with asymptotic behavior
|
|
|
Here, are the constant equilibria of (1.1): that is, . If , then is called a pulse (and a front otherwise). In this paper, we are focused primarily on fronts. We will always consider the case of : otherwise, the direction of motion can be reversed.
Traveling waves being solutions of (1.1) is an important subjection in dynamical systems, and many aspects of traveling waves have been studied in extensive works[11, 6]. One of the key issues of concern is whether a steady state is stable with respect to disturbances under initial conditions since this directly determines whether it can be observed in nature. Motivated by [9], in this paper, we establish a unified geometric approach for the stability analysis of traveling front solutions for (1.1).
Written in a moving frame, a traveling front solution of (1.1) can be regarded as a heteroclinic solution of the following equation:
|
|
|
(1.2) |
The stability analysis is directly related to the spectral information of the operator by linearizing (1.2) along , where ; note that the matrices are well-defined. Moreover, the previous description guarantees that there exists , such that
|
|
|
(1.3) |
For , denote by and the real and imaginary parts of , respectively. , , and . The pair denotes the -dimensional Euclidean space. Moreover, denotes the closure of the set .
Since a wave solution of (1.1) possesses a translation invariance property, it is said to be nondegenerate if zero is a simple eigenvalue of .
Definition 1.1.
A nondegenerate wave solution of (1.1) is spectrally stable if all the nonzero eigenvalues of are in .
In this paper, we study the following special case.
(H1).
Suppose that .
Thus, and are both stable equilibria of (1.1).
From now on, we focus on studying the eigenvalue problem
|
|
|
(1.4) |
Denoted by is the set of isolated eigenvalues with finite multiplicity, and is the essential spectrum of . It is known (Cf. Lemma 2.3) that . As was commonly performed in [9], we set to convert (1.4) to
|
|
|
(1.5) |
where , and there are well-defined matrices . Throughout this paper, the dots denote differentiation with respect to .
For any , denote by () the positive (negative) spectral space corresponding to the eigenvalues of having positive (negative) real parts.
We provide the next two conditions:
(H2).
, for all ,
(H2’).
, for all .
Letting , in the following remark, we will explain the relationship between (H(H1)) and (H(H2’)).
Remark 1.2.
(1) We claim that (H(H2’)) implies that (H(H1)). Let be the eigenvalue of with eigenvector . If , then we have that , which implies that . If , then is also the eigenvalue of with eigenvector , where and denote the complex conjugate of and , respectively. Now, let and : then, we have that
|
|
|
(1.6) |
Similarly, we have that
|
|
|
We thus determine that This implies that has a negative real part.
(2) We remark that (H(H2’)) cannot be derived from (H(H1)). We consider with the form and . It is easy to determine that , but .
(3) Supposing that , we claim that there exists a nonsingular matrix , such that
|
|
|
We only consider a special case, where the general case is analogous. Letting , suppose that and are generalized eigenvalues of and have the same algebraic multiplicity 2: then, for , there exists , such that
|
|
|
By a simple calculation, we have that
|
|
|
|
(1.7) |
|
|
|
|
(1.8) |
Following Definition 1.1, we focus our attention on : suppose that is the matrix solution of (1.5), such that .
We recall that the stable and unstable subspaces of (1.5) are
|
|
|
(1.9) |
Throughout this paper, we abbreviate and as and , respectively.
To realize the symplectic structure, we introduce the matrix
.
Since and , it follows that and . Therefore, is a complex structure on , and defines a syplectic form on . denotes the set of all Lagrangian subspaces of .
By invoking [13, Lemma 3.1], Remarks 2.4 and 2.6 imply that the subspaces and are both Lagrangian for . Moreover, those results can be obtained from the same discussion in [9, Theorem 2.3].
We denote by the space of all ordered pairs of continuous maps of Lagrangian subspaces equipped with the compact-open topology. Following the authors in [4], we are in the position to briefly recall the definition of the Maslov index for pairs of Lagrangian subspaces, which will be denoted throughout the paper by the symbol . Loosely speaking, given the pair ,
this index counts with signs and multiplicities the number of instants
that .
Definition 1.3.
The CLM-index is the unique integer valued function
|
|
|
satisfying Properties I-VI given in [4, Section 1].
For the sake of the reader, we list a couple of properties of the -index that we shall use throughout the paper.
-
•
(Reversal) Let . Denoting by the path traveled in the opposite direction, and by setting , we obtain
|
|
|
-
•
(Stratum homotopy relative to the ends) Given a continuous map
|
|
|
such that and are both constant, and then,
|
|
|
Moreover, one efficient way to study the Maslov index is via the crossing form introduced by [17] as follows.
Let be a smooth curve with . Let be a fixed Lagrangian complement of . For and small , define by . The form is independent of the choice of [17]. A crossing for is some for which intersects nontrivially. At each crossing, the crossing form is defined as
|
|
|
(1.10) |
A crossing is called regular if the crossing form is nondegenerate. For a quadratic form , we use the notation for its signature. We also write and for the positive and negative indices of inertia of , so that
|
|
|
From [20], if the path has only regular crossing with respect to , then we have that
|
|
|
(1.11) |
We recall the definition of Maslov index for a traveling pulse wave of (1.1) defined in [9]. The only requirement is that is large enough so that
|
|
|
(1.12) |
Moreover, since the transformation used in the conversion of (1.4) to (1.5) is different from that in [9], then the crossing form defined in the previous paper [9] differs from ours by a negative sign, and so, we provide [9, Definition 3.6] in the following form:
Definition 1.4.
[9, Definition 3.6]Let be a traveling pulse solution of (1.1), where the Maslov index of is given by
|
|
|
By using the Maslov index, this paper [9] provides a unified geometric treatment for the stability analysis of the traveling pulses solution and has been successfully employed in [11, 10] to obtain some elegant stability results. Motivated by [9], it is natural to consider the case of the traveling front solutions. We remark that Definition 1.4 is dependent on the condition (1.12). It is easy to check that this condition may be inefficient for the traveling front solutions, and we sidestep this difficulty by following [13, Definition 1.2] and define the Maslov index for the traveling waves as following.
Definition 1.5.
Let be a traveling wave of (1.1), and define the Maslov index of as
|
|
|
Remark 1.6.
If Definition 1.4 is well-defined, then it should be independent of the choice of . It was shown in [7] that this definition is independent of , as long as (1.12) is satisfied. However, for the traveling front solution, there may be an intersection point between and at . Thus, (1.12) may be inefficient. We sidestep this difficulty by following [13, Definition 1.2], and the following proposition shows that Definition 1.5 generalizes Definition 1.4 to the front case.
Proposition 1.7.
Letting be a traveling pulse solution of (1.1) and supposing that (H1) holds, then there exists , such that for all and
|
|
|
(1.13) |
Lemmas 2.10 and 2.5 show that the set of nonnegative, real eigenvalues of is bounded above,
and then, the spectrum of in consists of isolated eigenvalues of finite multiplicity (Cf. page 172 of
[2]), and so, it follows that the quantities
-
•
-
•
are both well-defined.
Now, we introduce a symplectic invariant called triple index (Cf. Definition 4.3) which
can be used to compute the Maslov index, and we denote by the triple index for any .
Theorem 1.8.
If (H(H1)) and (H(H2)) hold, then
|
|
|
(1.14) |
Remark 1.9.
Under the conditions (H1) and (H2), if (1.12) holds, then Definition 1.4 is well-defined for the traveling front solution. Using the notation of Maslov box in [9, Figure 4.1], by the same discussion in the proof of Proposition 1.7, it is easy to prove that the Maslov index contribution along the “bottom shelf” is equal to . Based on this discussion and Theorem 1.8, the distinction between the pulse and front is nontrivial, and we generalize [9, Theorem 4.1] to the front case.
Now, we present the central result of this paper.
Theorem 1.10.
If (H(H2’)) holds, then we have that
|
|
|
(1.15) |
Remark 1.11.
If (H(H2’)) holds, by Lemmas 3.4 and 3.5, we have that , and this implies that if (1.12) holds, then Definition 1.4 is well-defined. By invoking (3.11) and Lemma 3.6, (H(H2’)) provides a sufficient condition for the Maslov index contribution along the “bottom shelf” when is equal to . This can be easily verified for , and it remains true for . This is important for practitioners who want to use the Maslov index to prove stability. As shown in [11], an employed strategy can be used to compute for the doubly-diffusive FitzHugh-Nagumo system. Based on those, Definition 1.4 is more practical than Definition 1.5 if one wants to compute the Maslov index for this case. Moreover, from the same discussion in the proof of Proposition 1.7, under (H(H2’)), Definitions 1.4 and 1.5 are equivalent. The strategy employed in [11] may be valid for computing for the doubly diffusive FitzHugh-Nagumo system, such as to consider a traveling front solution established in [11, Theorem 2.2].
For the following FitzHugh-Nagumo equations
|
|
|
(1.16) |
where and with . If is large enough, then the equation has three solutions , and we remark that it is easy to check that and . Now, we consider a traveling front solution of the FitzHugh-Nagumo equation (1.16): its
existence has been established in [6], and we remark that such waves are obtained as local minimizers of an energy functional, and based on the variational characterization, authors [5] utilize the spectral flow to define and calculate a stability index. As a supplement, we provide a geometric insight into the stability of the traveling front solution for the FitzHugh-Nagumo equation (1.16).
Making a comparison with (1.1), then, (1.16) can be expressed as
|
|
|
(1.17) |
where , and
The eigenvalue problem
|
|
|
or its equivalent eigenvalue problem
|
|
|
(1.18) |
is studied to determine the stability of , where and .
Theorem 1.12.
Letting be a traveling front solution of the FitzHugh-Nagumo equation (1.16) satisfying the following asymptotic condition:
|
|
|
and
, then we have that
|
|
|
(1.19) |
2. Preliminary: an index formula
As is commonly done for the Sturm-Liouville operators [7, 14], in this section, by applying [13, Theorem 1], we derive an index formula (Cf. Proposition 2.12) for the Hamiltonian system (2.4), which plays a crucial role in obtaining the spectral information of .
The notation of spectral flow was introduced by Atiyah, Patodi and Singer in their study of index theory on manifolds with a boundary [3]. For the reader’s convenience, we provide a brief description of the basic properties of spectral flow.
Suppose that is a real separable Hilbert space, and denote by the space of all closed self-adjoint and Fredholm operators equipped with the gap topology. Let
be a continuous curve.
The counts the algebraic multiplicities of the spectral flow of across the line with some small positive number .
For each , there is an orthogonal splitting
|
|
|
where is the null space of , if and if .
There is an efficient way to compute the spectral flow through what are called
crossing
forms. Let be the orthogonal projector from to
. When
, we call
the instant a crossing instant. In this case, we defined the crossing
from
as
|
|
|
We call the crossing instant regular if the crossing form
is
nondegenerate. In this case we define the signature simply as
|
|
|
We assume that all of the crossings are regular. Then, the crossing instants are isolated (and, hence, those on a compact interval are finite in number), and the spectral flow is given by the following formula
|
|
|
(2.1) |
where , and denotes the set of
all crossings.
We list some basic properties of spectral flow:
Proposition 2.1.
-
(1)
If , then
|
|
|
-
(2)
If and denotes the set of
all crossings, then
|
|
|
Now, we return to the problem of stability analysis.
If there is not a Hamiltonian structure for by the presence of the term, we can circumvent this by considering instead the operator
|
|
|
(2.2) |
as is done in [12, 9].
We now proceed to the study of the following eigenvalue problem
|
|
|
(2.3) |
Setting and converts Equation (2.3) to the following Hamiltonian system
|
|
|
(2.4) |
where We define the self-adjoint operators
|
|
|
(2.5) |
where and . Here, is given in (1.3).
The following Proposition 2.2 shows that the Hamiltonian system (2.4) has close contact with the system (1.5).
Proposition 2.2.
Letting and , then if and only if .
It is well-known [16, Lemma 3.1.10] that the essential spectrum is given by
|
|
|
Following the same method as that in paper [11, Section 2], we here give the proof of the following Lemma for convenience of the reader.
Lemma 2.3.
Under the condition (H(H1)), we have that
-
(1)
,
-
(2)
if , then are hyperbolic.
Proof.
A simple calculation shows the eigenvalues of and , such that
|
|
|
(2.6) |
where is the eigenvalue of , and is the eigenvalue of . To prove that
(1) holds, we only need to show that and have no purely imaginary
eigenvalues if , which is equivalent to showing that and . From the formula we have that
|
|
|
(2.7) |
we note that , so Equation (2.7) implies that
|
|
|
(2.8) |
This calculation actually proves that has exactly eigenvalues of the positive real part and eigenvalues of the negative real part for . Additionally, . We label the eigenvalues of in order of increasing real part and observe that
|
|
|
(2.9) |
|
|
|
(2.10) |
From those, we complete the proof.
∎
Remark 2.4.
We note that : from Lemma 2.3, if (H(H1)) holds, then are both hyperbolic for all , and this, together with [14, Theorem 1], tells us that for each , the operator is a self-adjoint Fredholm operator, and in particular, it is possible to associate it to path
|
|
|
(2.11) |
the topological invariant: spectral flow.
The proof of Proposition 2.2.
If . By a simple calculation,
|
|
|
(2.12) |
From Lemma 2.3, must decay as fast as as , and we note that if , then and both exponentially decay to as , and hence, and . Then, : this together with Equation (2.12) implies that
.
Conversely, if , then by a simple calculation,
|
|
|
(2.13) |
We note that when , for each , invoking Lemma 2.5, it is easy to check that :
then, must decay at least as fast as as . We thus determine that and both exponentially decay as , and then, , and this together with Equation (2.13), implies that .
As a direct consequence, the following result holds.
Corollary 2.5.
Letting and , then if and only if .
Remark 2.6.
Under the condition (H(H1)), if satisfies (1.5) with , then a similar discussion in the proof of Proposition 2.2 guarantees that satisfies (2.4) with . Then, of the Hamiltonian system (2.4), so we can say that is also the stable space of the Hamiltonian system (2.4). By the same reasoning, is also the unstable space of the Hamiltonian system (2.4). Moreover, letting , a similar discussion in Proposition 2.2 shows that the following facts
and hold.
Under the condition (H(H1)), we determine that
|
|
|
(2.14) |
where the convergence is meant in the gap (norm) topology of the Lagrangian Grassmannian (Cf. [1] for further details).
Given , let with and with . The aim of the next part is to provide some sufficient condition for the coefficient of (2.11) to obtain the nondegeneracy.
Lemma 2.7.
Let
|
|
|
(2.15) |
and
|
|
|
(2.16) |
With given in (1.3), assuming that (H(H1)) holds, and , then the system
|
|
|
(2.17) |
has only the zero solution.
Proof.
Assuming that the system has a solution , then we have that
|
|
|
(2.18) |
Integrating by part, we obtain
|
|
|
(2.19) |
|
|
|
(2.20) |
Let and . It is easy to see that . Then, by using the second condition in the above boundary value problem, we obtain
|
|
|
(2.21) |
If , then we infer that
if and only if for . This concludes the proof.
Let us now consider the associated first order differential operators and of and . A similar result holds.
Lemma 2.8.
With given in (1.3), assuming that (H(H1)) holds, and , then the system
|
|
|
(2.22) |
has only the zero solution.
Lemma 2.9.
With given in (1.3), assuming that (H(H1)) holds, we have that
|
|
|
(2.23) |
Proof.
Let with and with . Then, the stable
subspace of the equation at is , the unstable
subspace of the equation at is , and there exists a linear bijection from the set of solutions
of the system
|
|
|
(2.24) |
with the subspace . By invoking once again Lemma 2.8 and Lemma 2.7, we
conclude that the initial value problem only admits the trivial solution for every . This concludes the proof.
∎
By setting for every and for every , the following result holds.
Lemma 2.10.
With given in (1.3), assuming that (H(H1)) holds, if , then and .
From Lemma 2.10, we determine that the following result holds.
Corollary 2.11.
With given in (1.3), assuming that (H(H1)) holds, and if , then we have that , where .
It is well-known that for each , the operator is closed and self-adjoint with dense domain in . As a byproduct of condition (H(H1)) and [14, Theorem 1], is also a Fredholm operator.
Finally, from [13, Theorem 1], we obtain that
|
|
|
|
(2.25) |
|
|
|
|
(2.26) |
As a direct consequence of Lemma 2.9 and Equation (2.25), we obtain the following result.
Proposition 2.12.
Under the previous notations and assuming that (H(H1)) holds, the following equation holds:
|
|
|
(2.27) |
3. The proof of the main result
The goal of this section is to provide a detailed proof of Theorem 1.8, Theorem 1.10 and Theorem 1.12. Before showing the proof, we start by analyzing the distribution of the eigenvalues of .
As in [8], the same discussion can be used here to obtain the distribution of eigenvalues of , which serves a crucial role of counting the number of nonnegative eigenvalues of via spectral flow. Let and be the orthogonal projections from to and , respectively. Define , and : in other words, can be decomposed as
|
|
|
(3.1) |
where and denotes the complex conjugate of . For a linear self-adjoint operator defined on a Hilbert space , denoted by , if for all , two linear operators and are denoted by if .
Lemma 3.1.
[8]
Supposing that and , then . The same assertion holds if and .
Proposition 3.2.
Under the condition (H(H1)), we have that
(1) if and , then ,
(2) if and , then .
Proof.
We only prove (1), while the other is analogous.
From Remark 2.6, we know that . Next, we prove that . Suppose that along the spectral flow, there is a crossing at for some and : that is,
|
|
|
(3.2) |
Letting and , we can rewrite Equation (3.2) as
|
|
|
(3.3) |
|
|
|
(3.4) |
Solving Equation (3.4), we obtain that , and this, together with Equation (3.3), obtains
|
|
|
(3.5) |
We note that for all . This indicates that the sign
of the crossing form has to be positive whenever a crossing occurs at . In view of Equation (2.1), we conclude from Lemma that
|
|
|
(3.6) |
For (2), a slightly modified argument shows that the sign of the crossing operator must be negative if a crossing occurs at , and then
|
|
|
(3.7) |
This completes the proof.
∎
The aim of the next part is to prove some transversal properties about some invariant subspaces that are useful in our proof.
Lemma 3.3.
Under the conditions (H(H1)) and (H(H2)), we have that
|
|
|
(3.8) |
.
Proof.
We provide the proof of in completely similar fashion. Let , and noting that is invariant under , then . From (H(H2)), a direct computation yields that
|
|
|
|
(3.9) |
|
|
|
|
(3.10) |
we determine that . This completes the proof.
∎
Now, following Lemma 4.4 and Equation (4.3), we have that
|
|
|
|
(3.11) |
|
|
|
|
(3.12) |
We recall that a Lagrangian frame for a Lagrangian subspace is an injective linear map whose image is (Cf. page 828 of [17]). Such a frame has the form
|
|
|
where and are both matrices and .
We introduce the following notations , , and , where are matrices, are matrices, are matrices, is the identity matrix, and is the identity matrix. Here, . For each , from Lemma 3.3, we can use notations and for the Lagrangian frames of and , respectively.
We now consider the following operator:
|
|
|
(3.13) |
and the associated second order operator and let be a solution of , where denotes the operator defined on the maximal domain . Then, the map provides a linear bijection from to .
We note that for each , there exists , such that .
Let with , where . A simple calculation shows that
|
|
|
|
(3.14) |
|
|
|
|
(3.15) |
|
|
|
|
(3.16) |
this equation, together with (H(H1)), (H(H2)) and (H(H2’)), shows the following Lemma:
Lemma 3.4.
With given in (1.3), the following results hold:
(1) if (H(H1)) and (H(H2)) hold, we have that is negative definite for all .
(2) if (H(H2’)) holds, we have that is negative definite for all .
Lemma 3.5.
With given in (1.3), the following results hold:
(1) if (H(H1)) and (H(H2)) hold, we have that is positive definite for all .
(2) if (H(H2’)) holds, we have that is positive definite for all .
Letting , then there is , such that , and can be rewritten as .
From a simple calculation, we have that
|
|
|
|
(3.17) |
|
|
|
|
(3.18) |
|
|
|
|
(3.19) |
|
|
|
|
(3.20) |
From Lemma 3.4, Lemma 3.5, Equation (3.17) and Equation (4.3), the following result holds.
Lemma 3.6.
With given in (1.3), the following results hold:
(1) if (H(H1)) and (H(H2)) hold, then we have that
|
|
|
(3.21) |
(2) if (H(H2’)) holds, then we have that
|
|
|
(3.22) |
Before finishing the preparation of our proof of our main results, we recall the definition of positive curve.
Definition 3.7.
[15]
Let
be a continuous curve. The curve is named a positive curve if is finite and
|
|
|
The Proof of Proposition 1.7.
For some , we construct the following homotopy Lagrangian path
|
|
|
(3.23) |
We point out that is constant for all and .
By the stratum homotopy invariance property of the Maslov index, we have that
|
|
|
|
(3.24) |
|
|
|
|
(3.25) |
We know that and as under the gap topology of the Lagrangian Grassmannian, so we can choose , such that for all , and the path has only regular crossing with respect to . Letting , we construct the following homotopy Lagrangian path:
|
|
|
By the stratum homotopy invariance, reversal property of Maslov index and Equation (1.11), we determine that
|
|
|
(3.26) |
|
|
|
(3.27) |
and this, together with Equation (3.24), completes the proof.
The Proof of Theorem 1.8.
We first prove that .
We start by introducing the continuous map
|
|
|
Let for . Then, for every , is a positive curve.
Let be a crossing instant for the path , meaning that , and let us consider the positive path . Thus, there exists , such that , which is equivalent to . Since is a Fredholm operator, then there exists , such that for every . By this argument, we determine that for every . Then, we determine that
|
|
|
(3.28) |
By the homotopy invariance of the spectral flow, we infer that
|
|
|
(3.29) |
and
|
|
|
(3.30) |
We observe that and are both positive curves. It follows that
|
|
|
|
(3.31) |
|
|
|
|
(3.32) |
from Equations (3.29), (3.30) and (3.31), we have that
|
|
|
|
(3.33) |
|
|
|
|
(3.34) |
the crossing instants are isolated, and those on a compact interval are finite in number. From Equation (3.33) and the path additivity of spectral flow, we determine that .
From Proposition 2.12 and Equation (3.11), we determine that
|
|
|
(3.35) |
Since (H(H1)) and (H(H2)) hold, from Lemma 3.6, Equation (3.35) and Proposition 2.1, then
|
|
|
(3.36) |
∎
The proof of Theorem 1.10.
From Remark 1.2, (H(H2’)) implies that (H(H1)) holds, and then, from Theorem 1.8 and Lemma 3.6, we have that
|
|
|
(3.37) |
∎
The proof of Theorem 1.12.
We note that , and then, , . Since , it is easy to see that . Moreover, if , then we have that , so Proposition 3.2 (1) holds, and then, we have that . Moreover, by a simple calculation
and the facts and , it is easy to check that the condition (H(H2’)) holds. Then, from Equation (3.35) and Lemma 3.6, we complete the proof.
∎
4. The triple and Hörmander index
Recently, Zhu et al., in the interesting paper [19], deeply investigated the Hörmander index, particularly its relation with respect to the so-called triple index in a slightly generalized (in fact, isotropic) setting. Given three isotropic subspaces and in , we define the quadratic form as follows:
|
|
|
(4.1) |
where for , and , . By invoking [19, Lemma 3.3], in the particular case in which are Lagrangian subspaces, we obtain
|
|
|
(4.2) |
By [19, Lemma 3.13], we are in position to define the triple index in terms of the quadratic form defined above.
Definition 4.1.
Let and be three Lagrangian subspaces of symplectic vector space . Then, the triple index of the triple is defined by
|
|
|
(4.3) |
where is the Morse positive index of a quadratic form Q.
Another closely related symplectic invariant is the so-called Hörmander index, which is particularly important for measuring the difference in the (relative) Maslov index computed with respect to two different Lagrangian subspaces (we refer the interested reader to the celebrated and beautiful paper [17] and the references therein).
Let be four Lagrangian subspaces and
be such that and .
Definition 4.2.
Letting be such that and , the Hörmander index is the integer defined by
|
|
|
|
(4.4) |
|
|
|
|
(4.5) |
Remark 4.3.
As a direct consequence of the fixed endpoints homotopy invariance of the -index, it is actually possible to prove that Definition 4.2 is well-posed, meaning that it is independent of the path joining the two Lagrangian subspaces . (Cf. [17] for further details).
Let us now be given four Lagrangian subspaces, namely of symplectic vector space . By [19, Theorem 1.1], the Hörmander index can be expressed in terms of the triple index as follows
|
|
|
(4.6) |
Lemma 4.4.
[14]
Let and
be two paths in with
, and assume that and
are both transversal to the (fixed) Lagrangian subspace . We then obtain
|
|
|