1. Introduction
In this paper, we will discuss a specific class of Fokker-Planck systems
that can be viewed as a special type of
generalized diffusion models in the framework of the energetic-variational
approach [24, 21, 17]. Such systems are determined by the kinematic transport of the
probability density function, the free energy functional and the dissipation
(entropy production), [2, 37].
In particular
we are interested in the nonlinear equations with non-homogeneous diffusion and mobility
in a finite bounded domain with periodic boundary conditions.
These Fokker-Planck type models
are motivated by the studies of the macroscopic behavior of the systems
that involve various fluctuations [35, 23, 15, 13, 11, 14, 27].
As in our previous work, we systematically studied such Fokker-Planck type
systems as a part of grain growth models in polycrystalline materials,
e.g. [4, 7, 3, 20].
One of the important goals of such study is to develop accurate mathematical
models that take into account critical
events, such as the grain
disappearance or nucleation, grain boundary disappearance, facet interchange, and
splitting of unstable junctions during
coarsening process of microstructure. For example, the recent model derived in
[20] under assumption of isothermal
thermodynamics can be viewed as a further
extension of a simplified 1D critical event model
studied in [5, 4, 7, 3]. In [20],
we have established the long time asymptotic results of the corresponding Fokker-Planck solutions, in terms of the
joint probability density function of misorientations (a difference in
the orientation between two neighboring grains that share a grain boundary) and triple
junctions (triple junctions are where three grain
boundaries meet), as well as the relation to the
marginal probability density of misorientations.
For an equilibrium configuration of a
boundary network, we obtained the explicit local algebraic
relationships, the generalized Herring Condition
formula, as well as a novel relationship that connects grain boundary energy density with the geometry of
the grain boundaries that share a triple junction.
The nonlinear Fokker-Planck
equations proposed in this work also appear as a part of our
current study of important case of non-isothermal
thermodynamics [33, 9, 38]. Such
Fokker-Planck systems
are derived with applications to
macroscopic models for grain boundary dynamics in polycrystalline
materials [19, 18, 6, 17].
Most existing mathematical analysis work of the Fokker-Planck models is
developed for the simplified linear cases only. This is especially true
for the well-known entropy methods developed for the asymptotic analysis of such
equations, e.g. [1, 28, 32, 12]. The
classical entropy methods
rely on the specific algebraic structures of the system, under convexity assumptions
of the potential functions and consider systems in unbounded domains.
In this paper, we study the generalized nonlinear Fokker-Planck models,
with inhomogeneous diffusion and/or mobility
parameters in a bounded domain with periodic boundary conditions.
This geometric constraint is relevant to our underlying grain boundary
applications, together with the non-convexity constraints of the
potential functions. Here we want to point out
that in the paper [16], we had overlooked this crucial
assumption in our mathematical analysis.
While the results there are valid, the mathematical analysis becomes
much more involved with less restrictive
assumptions on the models.
This is demonstrated in our current paper. Moreover, the mathematical
analysis results in
this paper are stronger than in [16] and in close agreement with the numerical
results in [16] in the absence of the convexity conditions.
The paper is organized as follows. Below, we formulate the nonlinear Fokker-Planck model with the inhomogeneous diffusion and
variable mobility parameters, introduce notations and some basic lemmas needed for the later sections. In Section 2, we first illustrate
large time asymptotic analysis for the homogeneous case. In this case, the equations become the usual linear Fokker-Planck model.
We employ the idea of the entropy method in terms of the velocity field of the
solution. In Sections 3-4, we extend the analysis to
the Fokker-Planck model with the inhomogeneous
diffusion without or with
variable mobility parameters respectively. Some conclusion remarks are
given in Section 5.
Let us start with the following Fokker-Planck type of equation
subject to the periodic boundary condition on a domain
(1) |
|
|
|
Here ,
are given positive
periodic functions with respect to and is a
given periodic function with respect to . The periodic boundary condition for
means,
(2) |
|
|
|
for
(3) |
|
|
|
and . In other words, can be smoothly
extended to a
function on the entire space with the condition
for , and , where
, with the 1 in the th place. Note that,
the periodic boundary condition for the function is equivalent
to the condition that is the function on the -dimensional
torus for . The periodic function is defined in the same way. The meaning of the periodic boundary condition for the Fokker-Planck equation can be seen in [34, §4.1, p.90].
The above equation can be viewed as a generalized diffusion in the
general framework of energetic variational approaches
[17, 24, 31, 39, 40].
One can see this by introducing a virtual velocity field ,
(4) |
|
|
|
and rewrite the (1) in a equivalent form involving the following
kinematic continuity equation (conservation of mass):
(5) |
|
|
|
The form of the first equation in (5) will
motivate us to extend various conventional methods established for various
fluids and diffusions to the nonlinear Fokker-Planck
model with inhomogeneous temperature parameter .
Next, from
(5) together with integration by parts and
with the periodic boundary condition, it is easy to obtain that,
(6) |
|
|
|
Therefore, if is a probability density function on , we have,
(7) |
|
|
|
Let be a free energy defined by,
(8) |
|
|
|
Hence, we can establish the energy law for
(5).
Hereafter, denotes the standard Euclidean vector norm.
Proposition 1.1.
Let be a solution of the periodic boundary value problem
(5), be the velocity vector
defined in (4), and let be a free energy defined
in (8). Then, for ,
(9) |
|
|
|
Proof.
Note that and are independent of , so direct
computation yields
(10) |
|
|
|
Using (1) together with periodic boundary
condition and integration by parts, we obtain
(11) |
|
|
|
From (4), hence
(12) |
|
|
|
Thus, (9) is proved by combining
(10), (11), and
(12).
∎
Throughout this paper, we assume the Hölder regularity with
for coefficients , , and initial
datum ,
(13) |
|
|
|
where
|
|
|
|
|
|
In this paper, we will consider the classical solutions of (1) defined below. In principle, they will be
smooth enough so that all the derivatives and integrations evolved in
the equations and the estimates will make sense (see
[17, 22, 29, 30]).
Definition 1.2.
A periodic function in space is a classical solution of
the problem (1) in ,
subject to the periodic boundary condition, if , for , and satisfies equation (1)
in a classical sense.
Next we will prescribe those constants in terms of . associated with various quantities.
This is important for the computations and estimates in the later sections.
-
(1)
The bounds for the initial data:
(14) |
|
|
|
-
(2)
The lower bound for the diffusion , such that
(15) |
|
|
|
-
(3)
The positive bounds for the mobility:
(16) |
|
|
|
-
(4)
The bound for the derivatives of the mobility:
(17) |
|
|
|
(18) |
|
|
|
-
(5)
The bound of the derivative of :
(19) |
|
|
|
-
(6)
The lower bound of the Hessian of :
(20) |
|
|
|
It is clear that these constants/parameters will also satisfy various
relations among themselves and they are also
associated with the original system.
Lemma 1.4.
For a positive function , the following is true:
(21) |
|
|
|
for and defined earlier and being a unit square/cube.
Proof.
For any , we first note that
(22) |
|
|
|
From (19) and being a unit square/cube, we
have
(23) |
|
|
|
By the triangle inequality and (22), we get
(24) |
|
|
|
Therefore, (21) is deduced by taking infimum with
respect to in (24).
∎
We look at an equilibrium solution of (1) in the form
(25) |
|
|
|
where the constant is determined as
(26) |
|
|
|
Here one can derive the relation between the constant
and the potential function .
Lemma 1.5.
Let be in (25). Then
(27) |
|
|
|
Proof.
By the mean value theorem and (26), there exists
such that
(28) |
|
|
|
Thus,
(29) |
|
|
|
hence . The estimate (27)
is easily deduced.
∎
From (27), it follows immediately that
Lemma 1.6.
Let be defined by (25). Then, for ,
(30) |
|
|
|
In [16, Proposition 1.6], we have derived the following maximum
principle which is crucial to our analysis.
Proposition 1.7.
Let the coefficients , , , and a positive
probability density function satisfy the strong positivity
(15), (16), (14) and the
Hölder regularity (13) for . Let
be a classical solution of (1), then the following holds:
(31) |
|
|
|
for , .
From Proposition 1.7, we have the following
logarithm estimates. In particular, we will show that the bound of the solution
is independent of the diffusion bound .
Corollary 1.8.
As in the same assumptions in Proposition
1.7,
there exists a positive constant which
depends only on , the initial datum (in terms of
), the gradient of (in
terms of ), and the bounds of the potential
such that
(32) |
|
|
|
Moreover, is independent to the diffusion bound .
Proof.
By (31), we have by (15) that,
(33) |
|
|
|
From here we have the following by using the assumption
(15):
(34) |
|
|
|
Using (21) and (30), we obtain,
(35) |
|
|
|
Since , we have
(36) |
|
|
|
Here, the constant depends only on
, , and
.
∎
In addition, we would need to use later the estimate (42) of
in terms of below,
where
(39) |
|
|
|
To obtain (42), we use the following Jensen’s inequality
[26, V.19], [36, Theorem 4.3].
Proposition 1.10 ([26, V.19]).
Let f be a function from to . Then is convex if and only if,
(40) |
|
|
|
whenever ,
, and .
Applying Jensen’s inequality (40) to convex function
, we have that,
(41) |
|
|
|
for . Using, in (41), we obtain,
(42) |
|
|
|
2. Homogeneous diffusion case
In this section, we consider the cases with both diffusion and mobility
being homogeneous. In this case, is a constant, and without loss of
generality, we choose and take .
Here, we study the following evolution equation with periodic boundary
conditions.
(43) |
|
|
|
The free energy associated with (43) will take the form :
(44) |
|
|
|
As we had presented in Proposition 1.1, the following energy law will hold
for any solution of (43):
(45) |
|
|
|
In this setting, equation (43) is a linear parabolic equation. One can
use various established methods and techniques to investigate the long-time asymptotic behavior for a
solution of (43). For instance, making the change
of variable
|
|
|
one may associate the original equation with a self-adjoint operator on (See
[20]).
In this paper, we plan to study the long time behavior of the solutions, especially the
exponential decay through the investigation of higher order time derivative of the free energy functional.
We want to point out that the current method is related to the entropy method that had been
developed previously for various Fokker-Planck type of equations in
unbounded domains. We consider here bounded domain and our approach takes the full advantage of the kinematic structures, such as looking at the velocity variable .
Theorem 2.1.
Given the potential and the domain in in
(43), for any , there exists
which depends on , the potential (in
terms of the lower bound of the Hessian defined in
(20) and
), the bounds of the initial data
, defined on
(14), and , such that if (15)
holds and,
(46) |
|
|
|
then, the following dissipation rate decays exponentially in time:
(47) |
|
|
|
To prove Theorem 2.1, we first recall the second derivative of
the free energy [16].
Proposition 2.4 ([16, Proposition 2.9]).
For the free energy defined in (44), and
be a solution of (43), we obtain:
(48) |
|
|
|
Since might be non-positive, our plan is to use the second term,
to control the right hand side of (48).
We first establish the following
inequality.
Lemma 2.5.
Let be a solution of (43), and let
be given as in (43).
There exists which depends only on , (in
terms of , defined in
(14)), and such that
(49) |
|
|
|
where
for ,
and for .
Proof.
By the Sobolev inequality ([25, Lemma 7.12,
7.16],[10, p.313]), there is
which depends only on such that,
|
|
|
for , where is the
integral mean of , namely
|
|
|
Here,
for ,
and for .
Since has the scalar potential, we can compute,
(50) |
|
|
|
Hence we obtain by taking that,
|
|
|
Note from (32) (see Remark 1.9) that for all and . This
inequality yields
|
|
|
Letting
, we obtain
(49).
∎
From here, we can show the following Poincare type inequality for the
velocity .
Lemma 2.8.
Let be a solution of
(43), and let be given as in
(43). There exists a constant
which depends only on , the norms
of the initial data (in terms of ,
defined on (14)), and
such that,
(51) |
|
|
|
Proof.
We show (51) for the case . First, we
use the Hölder inequality and (7) that
(52) |
|
|
|
where we select to satisfy
so we can employ the
Sobolev’s inequality (49). Next, we use
(49) to have that,
(53) |
|
|
|
Letting , we obtain
(51). Note that depends on
, and
so we can take
which depends only on , ,
and .
For the cases , (49) holds for any
. If, for instance, by taking , we will be able to use
the exact same estimates as in the case above to get the result.
∎
Now, we are in a position to derive the time derivative of the integral
.
Proposition 2.9.
Let be a solution of (43), and let
be given as in (43). For any , there
exists which depends only on ,
, , , defined
on (14), and such that
(54) |
|
|
|
Proof.
From the second derivative of the free energy
(48) and the lower bounds of
(20), we have
(55) |
|
|
|
Next, we use the Poincare type inequality (51)
that
(56) |
|
|
|
Together with the two inequalities, we obtain
(57) |
|
|
|
Choose sufficiently large such that
(58) |
|
|
|
Then, from the energy law (45), we have that,
(59) |
|
|
|
This finishes the proof of the Proposation.
∎
Finally, we are in position to prove the main result of this Section, Theorem 2.1.
Proof of Theorem 2.1.
Applying the Gronwall inequality to the differential inequality
(54), we obtain
(60) |
|
|
|
Using (46), we obtain
(47).
∎
In Section 2, in the case that and are constants, we
derive the exponential decay of the time derivative of the free energy
by construction of the differential inequality
(54). The
crucial idea is to use the Poincare type inequality
(51) in Lemma 2.8.
In the following sections, the Poincare type inequality will also play crucial roles in the study.
3. Inhomogeneous diffusion case
In this section, we consider the following
evolution equation with inhomogeneous diffusion while the mobility remaining a constant.
(61) |
|
|
|
with periodic boundary conditions. Without loss of generality, we take
. We choose a strictly positive periodic
function with the positive lower bound, ,
i.e., for .
Again, the direct formal computations show that the free energy and the basic energy law (9) take
the following specific forms,
(62) |
|
|
|
and
(63) |
|
|
|
As discussed in [16], the energy law (63)
with the free energy (62) carries all the physics of
the system, and the system, together with the kinematic assumption of
, will yield the equation (61) by the energetic variational
approach. In particular, we note that inhomogeneity is the
source of the nonlinearity in (61). Like what we had proved
in the previous section, Theorem 2.1, the following theorem
states that for any function with bounded second derivatives, we
can show exponential decay for the dissipation rate in the energy law
provided the diffusion coefficient is sufficiently
large. However, besides assumptions made in Theorem 2.1, we
will make additional assumptions on the gradients of and
, as well as assumption on the number of dimension .
Theorem 3.1.
Assume . For a fixed constant , there exist positive
constants and which
depend on dimension , the lower bound of Hessian of in
terms of defined in
(20), , the initial data
(in terms of , defined
on (14)), the gradient of (in terms of
defined in (19)),
, and
, such that if (15)
holds and
(64) |
|
|
|
then, for , we have the following result:
(65) |
|
|
|
for a positive constant
.
As in section 2, we have the following Sobolev-type inequality
for the velocity . Although the proof is the same as in Lemma
2.5, due to the fact that is a potential
gradient in this case, the constant , which is in
Lemma 3.2, will depend on defined in
(19).
Lemma 3.2.
Let be a solution of (61), and let
be given as in (61). There exists
which depends only on , (in terms of
, defined in
(14)), the gradient of (in terms of
defined in (19)), and
such that
(66) |
|
|
|
where for and
for .
With the extra dependence of the constants, the same proof in Lemma 2.8
will yield the following Poincare-type inequality for the velocity .
Lemma 3.3.
Let be a solution of (61), and let
be given as in (61). There exists a constant
which depends only on , (in terms of
, defined in
(14)), the gradient of (in terms of
defined in (19)), and
such that
(67) |
|
|
|
In the next lemma, we obtain the following interpolation inequality.
Lemma 3.5.
Let . Let be a solution of (61), and let
be given as in (61).
Then we have
(68) |
|
|
|
Proof.
The proof follows from the same as those in [16, Lemma 3.14]
(See also Lemma 4.7) together with Lemma
3.2.
∎
We next recall the second derivative of the free energy [16], which can be obtained by
direct computation from the original system (61):
Proposition 3.6 ([16, Proposition 3.13]).
Let be a solution of (61) and let be given as in
(61). Then the following is true:
(69) |
|
|
|
Below out of total 7 terms on the right-hand side of
(69), we first consider the 3rd,
4th, and 7th integral of the right-hand side of
(69). Notice the 2nd and 6th
terms are with the right positive sign, while the 5th term includes the
cubic order of .
Lemma 3.7.
Let be a solution of (61), and let be given as
in (61). Then,
we have
(70) |
|
|
|
Proof.
Note that the integrand of the 6th term of
(69) is non-negative, we need to
estimate the integrands of the 3rd, 4th, and 7th integrals of
(69).
We have by Cauchy’s inequality with a fixed
(See [22, p.662]) that,
(71) |
|
|
|
(72) |
|
|
|
and
(73) |
|
|
|
Here we used (15), (19),
(32), and (42). Thus, using
all inequalities above in (69),
we arrive at the estimate (70).
∎
Next, we compute the term with cubic order of in (70).
Lemma 3.8.
Let be a solution of (61), and let be given as
in (61). Assume (15).
Then, we have
(74) |
|
|
|
Proof.
By Using (15), (19), and
(32), we have
(75) |
|
|
|
Next, employing the interpolation inequality (68)
and we arrive at:
(76) |
|
|
|
Finally combining (75) with
(76) and using ,
we obtain the result (74).
∎
Next, we study the integrals involving and in
(69), by using the positive 2nd term.
We will show that with large diffusion bound , we can reduce the original
(69) into a specific form.
Proposition 3.9.
Let be a solution of (61), and let be given as
in (61). Assume the lower bound of the diffusion , is large enough such that the following condition holds:
(77) |
|
|
|
where the constant is the bound of the solution and
is the bound of the gradient of , as defined in
(32) and (19) in Section 1.
Then we obtain,
(78) |
|
|
|
Proof.
Note that the condition (77) yields the following inequalities:
(79) |
|
|
|
(80) |
|
|
|
and
(81) |
|
|
|
First, combining (70) and
(74), we obtain
(82) |
|
|
|
We consider the integral of in
(82). Using
(20), we get,
(83) |
|
|
|
Then applying
the assumption
(80), we have
(84) |
|
|
|
Next we consider the integral of in
(82).
Using (79), we
obtain,
(85) |
|
|
|
Thus, by (81) that
(86) |
|
|
|
Combining (85) and
(86), we arrive at
(87) |
|
|
|
Finally, use (79)
for the coefficient of the last term in the right-hand side of (82), we have
(88) |
|
|
|
With the fact that and are
independent of ,
combining (84),
(87), and
(88), we obtain
(78).
∎
Using Proposition 3.9
and with suitably large ,
we further reduce inequality the dissipation rate functional (78).
Lemma 3.11.
Let be a solution of (61), and let
be given as in (61). Then for a given ,
there exists a sufficiently large positive constant
which depends only on ,
, ,
from (20)(the
lower bound of the Hessian of ), ,
defined in (14)(the bounds of the
initial datum ), defined in
(19)(the bound of the gradient of ),
appeared in (66), and
such that if (15) holds, then,
(89) |
|
|
|
Proof.
By Lemma 3.3 together with (15),
(90) |
|
|
|
Note that depends only on ,
, defined in
(14), defined in
(19), and .
Thus, first we take large such that the
assumptions (79),
(80), and
(81) hold. Next,
for , take further sufficiently large
such that
(91) |
|
|
|
Then, we can use
(78)
and (90),
hence we obtain
(89).
∎
Next, we will recall the following Gronwall-type inequality from
[16].
Lemma 3.12 (Lemma 3.16 in [16]).
Let be positive constants, such that . Let
be a non-negative function that satisfies the
following differential inequality,
(92) |
|
|
|
If
(93) |
|
|
|
then, we obtain for ,
(94) |
|
|
|
This Gronwall-type inequality will allow us to show the exponential decay of
the dissipation rate of the free energy in (63),
the main result of this Section.
Proof of Theorem 3.1.
For any , using Lemma
3.11, take sufficiently
large positive number . Then with specifically
the following quantities,
(89) becomes
(92):
|
|
|
Thus, if
(95) |
|
|
|
then by Lemma 3.12
(96) |
|
|
|
Taking
and
(97) |
|
|
|
we will arrive at the conclusion of Theorem 3.1.
∎
In this Section, we had demonstrated the exponential decay of the time
derivative of the free energy in the case that is inhomogeneous
and the mobility is constant. In the next section, we will consider the
case of inhomogeneous diffusion and variable mobility .
4. Inhomogeneous diffusion case with variable mobility
This section will be devoted to the
following nonlinear Fokker-Planck equation with inhomogeneity in both
diffusion and mobility , which are bounded periodic
positive functions defined in a bounded domain in the Euclidean
space of -dimension.
(98) |
|
|
|
Again, without loss of generality, we take .
For the convenience of the readers, we recall (as defined in Section 1)
that the periodic function is bounded from below with the
constant , and the periodic function
is bounded both from below and above by the positive constants
and , namely
|
|
|
for any and .
The free energy and the basic energy law (9)
still takes the standard form:
(99) |
|
|
|
and the system satisfies the following energy dissipation law:
(100) |
|
|
|
We will first state the main result of this section: for a given system in (98),
with suitable conditions of the initial data and the mobility, under relatively mild inhomogeneity conditions,
one can find a diffusion that
is large enough, such that the system will convergence exponentially
to a equilibrium.
Theorem 4.1.
Assume . For a fixed constant ,
there exist positive constants
and
, which depend only on the given constant , , the potential (in terms of
Hessian bound
defined in (20),
, and ),
the bound of initial data , defined in (14), the bound of
defined in (19), and the bounds for the mobility ,
defined in (16),
such that if (15), (17),
(18) hold and,
(101) |
|
|
|
then for , the following is true,
(102) |
|
|
|
for a positive constant
.
Lemma 4.3.
Let be a natural number,
for ,
the positive constants being defined as the bounds of the mobility
(16) and the bound of the derivative of the mobility
(18) respectively.
Let be a solution of (98), and let
be given as in (98).
If
(103) |
|
|
|
then
there exists which depends
only on , (in terms of ,
defined in (16)), (in terms
of , defined on
(14)), the gradient of (in terms of
defined in (19)), and
such that,
(104) |
|
|
|
where for , and
for .
Proof.
First, using the definition of , the positive bounds for the mobility (16), we have,
(105) |
|
|
|
Then, we can use Lemma 3.2 that
(106) |
|
|
|
Using the definition of and in
the positive bounds for the mobility (16), we obtain
(107) |
|
|
|
Here
(108) |
|
|
|
Thus, we obtain by the Young inequality and the definition of
that,
(109) |
|
|
|
Therefore, we arrive at
|
|
|
Finally, we compute the coefficient of in the right hand
side. Using (103), the positive bounds for the mobility (16), and the bound for the derivatives of the mobility (18), we obtain
(110) |
|
|
|
hence we obtain (104) by taking
.
∎
With this result, next we will show that if the bounds for the gradient mobility
, defined in (18), is not too large,
then we can obtain the Poincare type inequality for .
Lemma 4.5.
Let be a solution of (98), and let
be given as in (98).
Assume the bound for the gradient mobility
, defined in (18) satisfies the following condition:
(111) |
|
|
|
where is the lower bound of appeared in (16), is appeared in
(104).
Then, there is a constant
which depends
only on , (in terms of ,
defined in (16)), (in terms
of , defined in
(14)), the gradient of (in terms of
defined in (19)), and
such that:
(112) |
|
|
|
Proof.
By the Hölder inequality and (7), we have
(113) |
|
|
|
We choose later and assume
(103). Then, by the Sobolev type inequality
(104), we obtain
(114) |
|
|
|
Now we choose as
(115) |
|
|
|
and take
Then, we obtain
(116) |
|
|
|
Plugging (115) into
(103), we get
(111). Taking
, we obtain
(112).
∎
In contrast to Lemma 3.2, the Sobolev-type of
inequality (104) has to include an extra quadratic term of in the
right hand side. Here we will re-derive the interpolation
inequality like that in Lemma 3.5.
Lemma 4.7.
Let and
let
, and let be
the constants defined in the bounds for the mobility
(16) and the bound of the derivative of the mobility
(18) respectively.
Let be a solution of (98), and let
be given as in (98).
Under the condition:
(117) |
|
|
|
we will have the following estimate:
(118) |
|
|
|
where is the constant defined in the Sobolev estimate
(104) in Lemma 4.3.
Proof.
We consider the case first. Let be constants
satisfying , and let be an exponent. Then, by the
Hölder’s inequality,
(119) |
|
|
|
where is the Hölder’s conjugate, namely, .
Next, we put a constraint, , in order to apply the
Sobolev type inequality (104) with .
Note that (117) guarantees the
assumption (103) in Lemma
4.3. Then (119) turns
into
(120) |
|
|
|
Next, we set other constraints, and
. The Young’s inequality implies,
(121) |
|
|
|
which yields:
(122) |
|
|
|
Now we can come back to examine the constraints. Since , , , and the properties of , imply that,
From these, one can deduce that , and in turns,
, and . Plugging these values in (122) and
we can obtain
(118) directly.
For the case , we can take , the same as in the case
. Then it is easy to verify that by taking and in
(122), one can obtain the inequality
(118).
∎
Next we recall the following higher order energy law of (98), which can be derived by
direct computations as we had done in [16].
Lemma 4.8 ([16, Proposition 4.15]).
Let be a solution of (98), and let be given as
in (98). Then, we have the following energy law,
(123) |
|
|
|
We will first derive the following inequality from the second derivative of
the free energy above (123).
Lemma 4.9.
Let be a solution of (98), and let be given as
in (98). Then,
we have
(124) |
|
|
|
where is the lower bound of
defined in (15), is the lower bound
of defined in (16), is
the upper bound for the estimate of the gradient of defined
in (18), is the upper bound for
the estimate of the gradient of defined in (19),
and is the bound for the estimate of
defined in (32).
Proof.
We start with the estimates of the integrands for the 3rd, 4th, and 7th
integrals of (123).
By Cauchy’s inequality with (See
[22, p.662]) and the definition of , we
will get
(125) |
|
|
|
(126) |
|
|
|
and
(127) |
|
|
|
Next we estimate the 10th, 11st, 12nd and 13rd terms of the
right-hand side of (123). Using the
Cauchy-Schwarz inequality that
(128) |
|
|
|
(129) |
|
|
|
(130) |
|
|
|
and
(131) |
|
|
|
Combining all inequalities above, ignoring the
integral of ,
and use them in (123), we arrive
at the reduced estimate (124).
∎
Next we consider the terms involving cubic order of in
(124).
Lemma 4.10.
Let be a solution of (98), and let be given as
in (98). Assume (117) and (15).
Then, we have,
(132) |
|
|
|
where is the lower bound of
defined in (15), is the upper bound
of defined in (16), is
the upper bound for the estimate of the gradient of defined
in (18), is the upper bound for
the estimate of the gradient of defined in (19),
is the bound for the estimate of defined in
(32), and is appeared in
(104).
Proof.
First, we compute the first term of the left hand side of
(132).
Using
(118) and
(15),
we compute
(133) |
|
|
|
Next, we consider the second term of the left hand side of
(132). Using
(118) and
(15) again that,
(134) |
|
|
|
Combining (133) and
(134), we obtain
(132).
∎
Now we are ready to demonstrate that under the assumptions
on suitable sufficient “smallness” conditions for the terms involving the gradient of the mobility , and , in terms of and ,
together with the assumption of being sufficiently
large (the lower bound of the diffusion),
one can deduce from (123) the following proposition.
Proposition 4.11.
Let be a smooth classical solution of
(98), and let be given as in (98). Assume
the lower bound of the diffusion , , is
large enough and the upper bounds of and , in
terms of and , are small
enough such that the following conditions hold:
(135) |
|
|
|
and
(136) |
|
|
|
Here, is the lower bound of
defined in (15), and
are the lower and the upper bounds of
defined in (16) respectively. The bound is
the bound for the estimate of the time derivative of defined in
(17), is the upper bound for the
estimate of the gradient of defined in (18),
is the upper bound for the estimate of the gradient
of defined in (19), is the
bound for the estimate of defined in
(32), and is appeared in
(104).
Then, we obtain,
(137) |
|
|
|
Proof.
Note that the conditions
(135),
, and
(136)
yield the following inequalities:
(138) |
|
|
|
(139) |
|
|
|
(140) |
|
|
|
(141) |
|
|
|
and
First, combining (124) and
(132), we obtain
(142) |
|
|
|
We consider the integral of in
(142). Using
(20) and (17), we
get
(143) |
|
|
|
Then applying the assumptions
(138) and
(139), we have
(144) |
|
|
|
Next we consider the integrals in (142) that contain term with . Note, using
(138), we obtain
(145) |
|
|
|
Here we use (21) and , then
(146) |
|
|
|
Thus, by the assumptions of
(140) and
(141), we have that
(147) |
|
|
|
Combining (145) and
(147), we arrive at
(148) |
|
|
|
Finally, we use (138)
to estimate the coefficient of the last term (the coefficient of the
integral cubed), and we have that,
(149) |
|
|
|
Combining (144),
(148), and
(149), we obtain
(137).
∎
Finally, the result in Proposition 4.11 will let us to the
desired differential inequality for the dissipation term and the proof of the main result Theorem 4.1.
Lemma 4.13.
Let be a solution of (98), and let be given as
in (98). Then for any , there exist a sufficiently
large positive constant and sufficiently small
positive constants which depend
only on , ,
, appeared in
(20)(the lower bound of the
Hessian of ), , defined
in (14)(the bounds of the initial datum ),
, defined in
(16)(the bounds of ), defined
in (19)(the bound of the gradient of ),
appeared in (104), and
such that
if (15), (17),
and
(18) hold,
then,
(150) |
|
|
|
Proof.
By Lemma 4.5,
(151) |
|
|
|
provided the relation (111):
With this assumption,
by (15) and (112), we have that,
(152) |
|
|
|
Note that depends only on ,
, , ,
, , and
, but is independent of
. Thus, for , take
sufficiently large such that
(153) |
|
|
|
Further, we take sufficiently large and sufficiently small
such that the assumptions
(138),
(139),
(140),
(141), and
(111) hold. Note that
from Corollary 1.8 and Lemma
4.5, and are
independent of , namely is bounded uniformly
with respect to . Then, we can use
(137) and
(154) |
|
|
|
Finally, using the bounds of in (16), we obtain
(150).
∎
Now we are ready to prove the main result of this Section.
Proof of Theorem 4.1.
For any , using Lemma
4.13, take sufficiently
large positive number and sufficiently small
positive numbers .
Then define,
|
|
|
and
together with
Thus, if
(155) |
|
|
|
then by Lemma 3.12
(156) |
|
|
|
Therefore, by taking
and
(157) |
|
|
|
we have finished the proof of Theorem 4.1.
∎