1 Introduction
Chemotaxis describes the oriented movement of species stimulated by uneven distribution of a chemical substance in the environment.
It is a significant mechanism accounting for abundant biological process/phenomenon, such as aggregation of bacteria [34, 44], slime mould formation [15], fish pigmentation [35], tumor angiogenesis [4, 7, 6], primitive streak formation [36], blood vessel formation [12], wound healing [39]. Mathematical models of chemotaxis were first proposed by Keller and Segel in their seminal works [19, 20, 21]. In this paper, we are concerned with the following chemotaxis model:
|
|
|
(1.3) |
where is a domain in with smooth boundary. System (1.3) was first advocated in [21] to describe the traveling band propagation of bacterial chemotaxis observed in the experiment of Adler [1, 2]. It later appeared in the work by Levine et al [23] to model the initiation of tumor angiogenesis, where
represents the density of vascular endothelial cells and denotes the concentration of signaling molecules vascular endothelial growth factor (VEGF). The parameters , are diffusion coefficients of the endothelial cells and the chemical VEGF respectively, is the chemotactic coefficient measuring the intensity of chemotaxis and is the chemical consumption rate by cells. In particular it was pointed out in [23] that the chemical diffusion process is far less important comparing to its interaction with endothelial cells and thus the diffusion coefficient could be small or negligible.
Despite of its biological significance, (1.3) is difficult to study mathematically due to the singularity of at . The well-known way to overcome this singularity was applying the following Cole-Hopf transformation
(cf. [22, 31]):
|
|
|
(1.4) |
to transform (1.3) into a system of conservation laws:
|
|
|
(1.5) |
The transformed system (1.5) attracts extensive attentions and numerous interesting results have been developed. We briefly recall these results by the dimension of spaces. In the one dimensional case, the global well-posedness along with large time behavior of solutions was investigated when in [13, 25] with and in [37, 33] with .
When , authors in [54, 28] obtained the unique global solution under Neumann-Dirichlet boundary conditions for , and the result was later extended to the case in [49, 43]. The problem with is also globally well-posed [26] with Dirichlet-Dirichlet boundary conditions. Furthermore, the existence and stability of traveling wave solutions were studied in [18, 29, 30, 31, 32, 27, 3]. However to the best of our knowledge, except when it is associated with radially symmetric initial data, the known well-posedness results of problem (1.5) in the multi-dimension are merely confined to local large and global small solutions, cf. [24, 14, 8, 38, 45, 48] for details when () and [28, 41] when () is bounded.
If the initial data are radially symmetric and (), Winkler [50] recently proved that (1.3) with subject to Neumann boundary conditions admits a global generalized (weak) solution which is radially symmetric and smooth away from the origin .
In addition to the above well-posedness results, the asymptotic behavior of solutions as is a particularly relevant issue
(mentioned in [23] that could be small/negligible) and has been studied in several circumstances for equation (1.5) and thus for the original equation (1.3) via transformation (1.4). For illustration, denote by and the solutions of (1.5) with and respectively (when we shall use the notation instead of since it is a scalar). First in unbounded domains, it has been shown that both traveling wave solutions (cf. [47]) in and the global small-data solution of the Cauchy problem (cf. [48, 38]) in are uniformly convergent in , namely converge to in -norm as .
With , the solutions still converge (cf. [49]) as when (1.5) is endowed with the following mixed homogeneous Neumann-Dirichlet boundary conditions
|
|
|
However if both of and are subject to the Dirichlet boundary conditions, one can not preassign the boundary value for when since it is intrinsically determined by the second equation of (1.5) as . Thus the plausible Dirichlet boundary conditions should be prescribed as:
|
|
|
(1.6) |
where are constants.
In this case, if the boundary values of with and do not match, then the solution component would diverge near the end points as and this phenomenon is termed as the boundary layer effect, which has been an important topic in the fluid mechanics [42] when investigating the inviscid limit of the Navier-Stokes equations near a boundary and has attracted extensive studies (cf. [9, 10, 11, 17, 46, 52, 53]) since the pioneering work [40] by Prandtl in 1904. In particular, this boundary layer effect for problem (1.5)-(1.6) has been recently numerically verified in [26] and rigorously proved in [16].
Enlightened by these results, it is natural to expect that (1.5) in multi-dimension () possesses boundary layer solutions as well when prescribing appropriate Dirichlet boundary conditions. In particular we aim to investigate this issue for its radial solutions in the present paper. To this end, we first rewrite (1.3) in its radially symmetric form by assuming that the solutions are radially symmetric, depending only on the radial variable and time variable . In a domain bounded by two concentric sphere, i.e. , (1.3) reads as
|
|
|
(1.10) |
where have been assumed without loss of generality.
Similar as deriving (1.5) from (1.3), we apply the following Cole-Hopf type transformation
|
|
|
(1.11) |
which turns (1.10) into
|
|
|
(1.15) |
Similar to (1.6), the Dirichlet boundary conditions for (1.15) are prescribed as
|
|
|
(1.16) |
In this paper, we shall investigate the asymptotic behavior of solutions to (1.15)-(1.16) as for (if , it coincides with the one-dimensional model (1.5)-(1.6) which has been studied in [16] as aforementioned). In particular, the solution component is proved to have a boundary layer due to the mismatch of its boundary values as (see Theorem 2.2).
2 Main results
To study the boundary layer effect, we first present the global well-posedness and regularity estimates for solutions of (1.15)-(1.16) with in Theorem 2.1. By these estimates, we then state the main result on the convergence for and boundary layer formation by in Theorem 2.2. Finally, the result is converted to the original chemotaxis model (1.10) via (1.11). We begin with introducing some notations.
Notations. Without loss of generality, we assume since the zero diffusion limit as is our main concern. Throughout this paper, unless specified, we use to denote a generic positive constant which is independent of and dependent on . In contrast, denotes a generic constant independent of and . For simplicity, represents with , denotes with and stands for . Moreover, if for fixed , we use to denote .
The first result is on the global well-posedness of (1.15)-(1.16) with .
Theorem 2.1.
Assume that with satisfy the compatible conditions
.
Then the initial-boundary value problem (1.15)-(1.16) with has a unique solution such that the following estimates hold true.
(i) If , there is a constant independent of such that
|
|
|
(2.1) |
Moreover,
|
|
|
(2.2) |
(ii) If , for any , there exists a constant depending on such that
|
|
|
(2.3) |
We proceed to recall the definition of boundary layers (BLs) following the convention of [10, 11].
Definition 2.1.
Denote by and the solution of (1.15)-(1.16) with and , respectively. If there exists a non-negative function satisfying as such that
|
|
|
we say that the initial-boundary value problem (1.15)-(1.16) has a boundary layer solution as and is called a boundary layer thickness (BL-thickness).
Our main result is as follows.
Theorem 2.2.
Suppose that with satisfy the compatible conditions and . Let be the solution obtained in Theorem 2.1. For any , we denote
|
|
|
where the function is defined in (4.9) by , and the constant (given in (4.19)) depends only on and .
Then (1.15)-(1.16) with admits a unique solution . Furthermore, any function satisfying
|
|
|
(2.4) |
is a BL-thickness of (1.15)-(1.16) such that
|
|
|
(2.5) |
|
|
|
(2.6) |
Moreover,
|
|
|
(2.7) |
if and only if
|
|
|
(2.8) |
By employing transformation (1.11), we next convert the above results for (1.15)-(1.16) to the pre-transformed chemotaxis model (1.10). The counterpart of the original model reads as follows:
|
|
|
(2.9) |
Proposition 2.1.
Assume and . Suppose that the assumptions in Theorem 2.2 hold with . Let . Then (2.9) with admits a unique solution such that
|
|
|
(2.10) |
Moreover, the gradient of has a boundary layer effect as , that is
|
|
|
(2.11) |
with the function defined (2.4)
and the following estimate holds
|
|
|
(2.12) |
if and only if (2.8) is true.
At the end of this section, we briefly introduce the main ideas used in the paper. Although the system (1.15)-(1.16) with is in a similar form to its counterpart with for which the vanishing diffusion limit has been studied in [16] based on a -independent estimate for solutions with , the methods used there can not be applied to study the present problem since when the system (1.15)-(1.16) with lacks an energy-like structure or a Lyapunov function to provide a preliminary estimate uniformly in . Moreover, one can not use the estimates derived in [50] for the present problem either since those estimates depend on . The difficulty in our analysis consists in deriving the -convergence estimates in (2.5) and (2.6) without any uniform-in- priori bounds on solutions . Inspired by the works [5, 51], this will be achieved in section 4 by regarding with small as a perturbation of and then estimating their difference by the method of energy estimates and a new Gronwall’s type inequality (see Lemma 4.1) on ODEs. The proof of Theorem 2.1 is standard and will be given in section 3.
3 Proof of Theorem 2.1
This section is to prove Theorem 2.1 based on the following lemmas where the a priori estimates on solution of (1.15)-(1.16) with are derived by the energy method. We set off by rewriting (1.15)-(1.16) with as follows:
|
|
|
(3.5) |
Lemma 3.1.
Suppose the assumptions in Theorem 2.1 hold and . Then there exists a positive constant independent of such that
|
|
|
(3.6) |
and
|
|
|
(3.7) |
Proof.
Taking the inner products of the first and second equation of (3.5) with and respectively, we then add the results and use integration by parts to get
|
|
|
which gives rise to (3.6) upon integration over . To prove (3.7), we denote and find from (3.5) that satisfies
|
|
|
(3.12) |
Multiplying the first and second equation of (3.12) by and , respectively. Adding the results gives
|
|
|
(3.13) |
Note that can be estimated as follows
|
|
|
Then substituting the above estimate into (3.13) and integrating the result over we have
|
|
|
which, along with (3.6) and the fact
|
|
|
implies (3.7). The proof is completed.
We proceed to derive higher regularity properties for the solution of (3.12).
Lemma 3.2.
Suppose the assumptions in Theorem 2.1 hold and . Let be the solution of (3.12). Then there is a constant independent of such that
|
|
|
(3.14) |
Proof.
We multiply the second equation of (3.12) with and differentiate the resulting equation with respect to . Then from the first equation of (3.12) we obtain
|
|
|
(3.15) |
Taking the inner product of against to get
|
|
|
(3.16) |
We may rewrite as
|
|
|
where can be reorganized as
|
|
|
and can be estimated by (3.15) and the Poincaré inequality as
|
|
|
Hence
|
|
|
To estimate , we first note that for fixed if satisfies it follows that which leads to
|
|
|
(3.17) |
thanks to the Poincaré inequality .
Then we deduce from (3.17) and the Sobolev embedding inequality that
|
|
|
(3.18) |
Substituting the above estimates for and into (3.16), one derives
|
|
|
Then applying Gronwall’s inequality to the above result and using Lemma 3.1, we conclude that
|
|
|
(3.19) |
We proceed to estimate by multiplying the first equation of (3.12) with in and derive
|
|
|
(3.20) |
By similar arguments as deriving (3.18), we estimate as
|
|
|
and by the
Cauchy-Schwarz inequality, is estimated as
|
|
|
Then feeding (3.20) on the above estimates for and , we have
|
|
|
(3.21) |
Integrating (3.21) over and using (3.7) and (3.19), one arrives at
|
|
|
which, in conjunction with (3.19) gives (3.14). The proof is completed.
Lemma 3.3.
Suppose that the assumptions in Theorem 2.1 hold and . Then there exists a constant independent of such that
|
|
|
(3.22) |
and
|
|
|
(3.23) |
Proof.
Differentiating the first equation of (3.12) with respect to and multiplying the result with , we get upon integration by parts that
|
|
|
(3.24) |
By (3.17) and the second equation of (3.12) we have that
|
|
|
We use again the second equation of (3.12) and Cauchy-Schwarz inequality to get
|
|
|
Substituting the above estimates for - into (3.24), then integrating the results over and using Lemma 3.1 along with Lemma 3.2, we conclude that
|
|
|
(3.25) |
We proceed to estimating the remaining part
in (3.22). Differentiating (3.15) with respect to
and multiplying the resulting equation with we get
|
|
|
(3.26) |
To estimate of ,
we note for with fixed , it follows that
|
|
|
(3.27) |
Then from Cauchy-Schwarz inequality and (3.27) one derives
|
|
|
To bound we first estimate by the first equation of (3.12) as follows:
|
|
|
(3.28) |
where (3.17) and Lemma 3.1 - Lemma 3.2 have been used. Then (3.28) along with (3.27) and (3.7) implies that
|
|
|
(3.29) |
where the constant depends on and . Noting that , one deduces by (3.17) and the Sobolev embedding inequality that
|
|
|
We feed (3.26) on the above estimates for - then apply Gronwall’s inequality, Lemma 3.1 - Lemma 3.2, (3.25) and (3.29) to the result to find
|
|
|
which, along with (3.25) yields (3.22). We next prove (3.23).
By similar arguments as deriving (3.28) one gets
|
|
|
(3.30) |
where (3.25) and Lemma 3.1 - Lemma 3.2 have been used. We differentiate (3.15) with respect to and conclude that
|
|
|
(3.31) |
where we have used (3.22), Lemma 3.1 and Lemma 3.2. Finally collecting (3.28), (3.30) and (3.31) we derive (3.23). The proof is finished.
We are now in the position to prove Theorem 2.1 by the above Lemma 3.1 - Lemma 3.3.
Proof of Theorem 2.1.
We first prove Part (i) of Theorem 2.1. By Lemma 3.1 and (3.27), one derives
|
|
|
(3.32) |
where the constant depends on and and the Poincaré inequality has been used.
On the other hand, for with fixed we have
|
|
|
(3.33) |
Then it follows from Lemma 3.2, Lemma 3.1, (3.27) and (3.33) that
|
|
|
(3.34) |
Similarly, it follows from Lemma 3.3 and (3.33) that
|
|
|
(3.35) |
Thus collecting (3.32), (3.34) and (3.35) we derive the desired a priori estimate (2.1), which along with the fixed point theorem implies the existence of solution in .
We next prove (2.2). Integrating (3.21) over with respect to , then using Lemma 3.1 and Lemma 3.2, we have
|
|
|
which, along with (3.7) implies that
Hence, it follows that
|
|
|
which, along with the Gagliardo-Nirenberg inequality and (3.7), gives (2.2).
Part (i) of Theorem 2.1 is thus proved.
We proceed to prove Part (ii). When , for one can easily deduce the a priori estimates (2.3) by the standard energy method that bootstraps the regularity of the solution from to . We omit this procedure for simplicity and refer readers to [26] for details. Then the existence of solution follows from (2.3) and the fixed point theorem. The proof is finished.
4 Proof of Theorem 2.2 and Proposition 2.1.
Let and be the solutions of (1.15)-(1.16) corresponding to and respectively. Then the initial-boundary value problem for their differences , reads:
|
|
|
(4.1) |
To prove Theorem 2.2 we shall invoke an elementary result (see Lemma 4.1) on an ordinary differential equation (ODE) and a series of lemmas on the a priori estimates for solutions of (4.1). In particular, the -estimate for solution and higher regularity estimates for the solution component will be established in Lemma 4.2 - Lemma 4.5, and Lemma 4.6 will give a weighted -estimate for the derivative of .
We proceed to prove the following Lemma, which gives an upper bound for the solution of an ODE involving a small parameter . It extends a result in [5, 51] with to any integer .
Lemma 4.1.
Let be an integer and . Let be a constant independent of and , be two continuous functions on . Consider the ODE
|
|
|
(4.2) |
If we set
|
|
|
(4.3) |
with .
Then for , any solution of (4.2) satisfies
|
|
|
(4.4) |
Proof.
Let . Then (4.2) can be rewritten as
|
|
|
Noting that thanks to ,
we deduce that
|
|
|
(4.5) |
For later use, we define
|
|
|
(4.6) |
Now dividing both sides of (4.5) by , it follows that
|
|
|
Then noting due to its definition (4.6), we deduce from the above inequality that
|
|
|
|
|
which integrated over with , yields
|
|
|
(4.7) |
where we have used the fact , thanks to the definition of .
We shall prove that
|
|
|
(4.8) |
of which the proof is split into three cases by the value of .
Case 1, when , it follows from the definition of in (4.3) that
|
|
|
Case 2, when , we have by using (4.3) again that
|
|
|
Case 3, when , one immediately get
|
|
|
Hence combining the above Case 1 - Case 3, we conclude that (4.8) holds true and it follows from (4.8) and (4.7) that
|
|
|
thus
|
|
|
which, along with (4.6) and the definition of , yields the desired estimate (4.4).
The proof is finished.
In the sequel, for convenience we denote
|
|
|
|
|
|
(4.9) |
The following lemma gives the -estimate for the solution of problem (4.1).
Lemma 4.2.
Let . Then there exists a constant independent of and , such that
|
|
|
(4.10) |
Proof.
Testing the first equation of (4.1) with in and using integration by parts, we get
|
|
|
The estimate of follows from (3.17) and (3.27):
|
|
|
On the other hand, the Sobolev embedding inequality and Cauchy-Schwarz inequality entail that
|
|
|
Collecting the above estimates for and , we conclude that
|
|
|
(4.11) |
We proceed by taking the inner product of the second equation of (4.1) with and get
|
|
|
First Sobolev embedding inequality and (3.27) yield
|
|
|
It follows from Cauchy-Schwarz inequality and (3.27) that
|
|
|
Moreover Sobolev embedding inequality, Cauchy-Schwarz inequality and (3.27) lead to
|
|
|
Collecting the above estimates for - and recalling that , we end up with
|
|
|
which, adding to (4.11) gives the desired estimate (4.10). The proof is completed.
We turn to estimate the derivative of and the boundary term in (4.10).
Lemma 4.3.
Let . Then there exists a constant independent of and , such that
|
|
|
(4.12) |
and
|
|
|
(4.13) |
Proof.
Taking the inner product of the second equation of (4.1) with , then using integration by parts we have
|
|
|
We first employ Sobolev embedding inequality and (3.27) to deduce that
|
|
|
and that
|
|
|
Moreover Cauchy-Schwarz inequality and (3.27) entail that
|
|
|
Then (4.12) follows from the above estimates on -. It remains to prove (4.13).
By the definition of and Gagliardo-Nirenberg interpolation inequality, one deduces that
|
|
|
(4.14) |
where is a small constant to be determined. By a similar argument as deriving (4.14) and the second equation of (1.15) with , we further get
|
|
|
(4.15) |
To bound the term in (4.14) and (4.15), we use the second equation of (4.1), Sobolev embedding inequality and (3.27) and derive
|
|
|
(4.16) |
where we have used the notation to distinguish it from the constant in (4.14) and (4.15).
Finally feeding (4.14) and (4.15) on (4.16) then adding the results, we obtain (4.13) by taking small enough such that and by using . The proof is completed.
We next apply Lemma 4.1 to the combination of Lemma 4.2 and Lemma 4.3 to obtain the following result.
Lemma 4.4.
Let and with defined in Theorem 2.2. Then there exists a constant independent of , depending on such that
|
|
|
(4.17) |
and
|
|
|
(4.18) |
Proof.
We first add (4.12) and (4.13) to (4.10) and find
|
|
|
(4.19) |
where has been used. Then we apply lemma 4.1 to (4.19) by taking and to conclude for that
|
|
|
(4.20) |
where (3.27) has been used.
Then we integrate (4.19) over (0,T) and applying (4.20) to the result to deduce that
|
|
|
where the constant depending on and is finite thanks to Theorem 2.1. We thus derive (4.17) and proceed to prove (4.18). Indeed, it follows from the second equation of (4.1), Sobolev embedding inequality and (4.17) that
|
|
|
We thus derive (4.18) and the proof is completed.
Higher regularity estimates for the solution component is given in the following lemma.
Lemma 4.5.
Suppose and . Then there is a constant independent of , depending on such that
|
|
|
(4.21) |
Proof.
We first take the inner product of the first equation of (4.1) with and use integration by parts to get
|
|
|
(4.22) |
Then differentiating the first equation of (4.1) with respect to and multiplying the resulting equation with in , we derive
|
|
|
The estimate for follows from (3.17) and (3.27):
|
|
|
Sobolev embedding inequality and (3.27) entail that
|
|
|
Then collecting the above estimates for and , one derives
|
|
|
(4.23) |
where we have used inequalities and , thanks to the first and second equations of (3.5).
Finally by adding (4.23) to (4.22) and applying Gronwall’s inequality to the result, and then using (4.17) we obtain (4.21). The proof is completed.
We turn to establish a weighted -estimate (enlightened by [17]) on the derivative of .
Lemma 4.6.
For and , there is a constant independent of , depending on such that
|
|
|
Proof.
Taking the inner product of the second equation of (4.1) with and using integration by parts, one gets
|
|
|
(4.24) |
We next estimate -. Indeed Cauchy-Schwarz inequality, Sobolev embedding inequality and (3.27) yield
|
|
|
and
|
|
|
For the term , we use integration by parts and the first equation of (4.1) to get
|
|
|
We proceed to estimate -.
First it follows from Cauchy-Schwarz inequality that
|
|
|
Moveover, we use Cauchy-Schwarz inequality and apply (3.17) to and to derive
|
|
|
The estimate for follows from the Sobolev embedding inequality, (3.17) and Cauchy-Schwarz inequality:
|
|
|
We thus conclude from the above estimates for - that
|
|
|
Substituting the above estimates for - into (4.24), then applying Gronwall’s inequality, Lemma 4.4, Lemma 4.5 and Theorem 2.1 to the result, we obtain the desired estimate and the proof is finished.
We next prove Theorem 2.2 by the results derived in Lemma 4.4 - Lemma 4.6.
Proof of Theorem 2.2.
By Lemma 4.4, Lemma 4.5 and Sobolev embedding inequality, we deduce that
|
|
|
which gives (2.5). Clearly holds for and , thus it follows from Lemma 4.6 that
|
|
|
which, along with Lemma 4.4 and Gagliardo-Nirenberg inequality entails that
|
|
|
provided .
Hence we derive (2.6) and we next prove the equivalence between (2.7) and (2.8). We first prove that (2.8) implies (2.7). Assume for some . Then integrating the second equation of (1.15) with over along with compatible condition gives
|
|
|
(4.25) |
We thus have
|
|
|
Similar arguments lead to (2.7) when assuming that
for some . We thus proved (2.7) provided (2.8). The proof of that indicates will follow from the argument of contradiction. Indeed, if we assume that (2.7) holds and the opposite of (2.8) holds, that is
|
|
|
which, along with the second equation of (1.15) with leads to
|
|
|
where the compatible conditions have been used.
Thus and and the terms , in (4.10) and (4.12) would vanish. Then by similar arguments as deriving (4.20), we conclude that
|
|
|
which, along with Sobolev embedding inequality gives rise to
|
|
|
which, contradicts with (2.7), thus (2.7) implies (2.8).
The proof is completed.
We next convert the result of Theorem 2.2 to the initial-boundary value problem (2.9) for the original chemotaxis model to prove Proposition 2.1.
Proof of Proposition 2.1.
Let
|
|
|
(4.26) |
where and are the solutions of problem (1.15)-(1.16) with and respectively. It is easy to check that and with and defined (4.26) are the unique solutions of (2.9) corresponding to and , respectively.
The first inequality in (2.10) follows directly from the Sobolev embedding inequality, Lemma 4.4 and Lemma 4.5 as following:
|
|
|
To prove the second inequality in (2.10), we deduce from (4.26) that
|
|
|
|
(4.27) |
|
|
|
|
|
|
|
|
where we have denoted by for convenience.
For , it follows from Lemma 4.4 and the Sobolev embedding inequality that
|
|
|
|
(4.28) |
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
for some constant independent of (depending on ) and the assumption has been used in the last inequality. Then we apply Taylor expansion to and using (4.28) to conclude that
|
|
|
(4.29) |
where the assumption has been used in the second inequality and the constant is independent of .
On the other hand, by the assumptions and in Proposition 2.1 we derive that for some positive constant , which along with the fact leads to
|
|
|
(4.30) |
Moreover, from Theorem 2.1 we know that
|
|
|
(4.31) |
where the constant depending on . Thus we deduce from the second equality of (4.26), (4.30) and (4.31) that
|
|
|
(4.32) |
with .
Hence, it follows from (4.27), (4.29) and (4.32) that
|
|
|
(4.33) |
where is independent of .
We thus derive the second inequality in (2.10) and proceed to prove (2.11).
It follows from transformation (1.11) that
|
|
|
(4.34) |
which, in conjunction with (2.6) and (2.10) leads to
|
|
|
where has been used thanks to as . We thus derived (2.11). To prove the equivalence between (2.12) and (2.8), we first derives two positive constants and independent of such that
|
|
|
(4.35) |
by choosing small enough such that in (4.33) and using (4.32).
With (4.35) in hand, we next prove the equivalence between (2.12) and (2.8).
First, it follows from (4.34), (2.10) and (4.35) that
|
|
|
(4.36) |
Dividing (4.34) by and applying a similar argument as deriving (4.36), then using (4.35) and (2.10), one deduces that
|
|
|
which, in conjunction with (4.36) indicates the equivalence between (2.12) and (2.7). Then we conclude that (2.12) is equivalent to (2.8) by using Theorem 2.2. The proof is completed.