1 Introduction
Overview of maximal regularity results in flat domains.
Let be a flat, polyhedral of dimensions or smooth domain of arbitrary dimensions, and consider the initial and boundary value problem for a linear parabolic partial differential equation (PDE)
|
|
|
(1.4) |
where are real-valued bounded measurable functions satisfying the following uniform ellipticity condition for some constant :
|
|
|
(1.5) |
Under condition (1.5), the elliptic partial differential operator generates a bounded analytic semigroup on , , and the solution of (1.4) possesses the following maximal -regularity in :
|
|
|
(1.6) |
which is an important mathematical tool in studying the well-posedness and regularity of solutions of nonlinear parabolic PDEs; see [4, 9, 36, 27, 39, 41].
Analogously, denoting by the finite element approximation of the elliptic operator on a finite element subspace , i.e.,
|
|
|
(1.7) |
it is known that the semi-discrete finite element solutions given by
|
|
|
(1.10) |
satisfies the following spatially discrete -uniform maximal -regularity [31, 35] (with a constant independent of the mesh size ):
|
|
|
|
(1.11) |
which has applications in numerical analysis for semilinear parabolic equations with strong nonlinearities [18], and quasi-linear parabolic equations with nonsmooth coefficients [34]. The spatially discrete maximal -regularity results were firstly proved in smooth domains with Neumann boundary condition [17, 31], and then extended to polyhedral domains [35, 32] with the Dirichlet boundary condition. The discrete maximal -regularity is also closely related (in the techniques of proof) to the maximum-norm stability of finite element solutions of parabolic equations [29, 40, 42, 43, 47]:
|
|
|
(1.12) |
The extension of maximal -regularity to the time-discrete setting was established for different time discretization methods, including the backward Euler method [5], discontinuous Galerkin method [30], -schemes [21], and A-stable multistep and Runge-Kutta methods [24]. All these methods are A-stable. The maximal -regularity of A-stable backward differentiation formulae (BDF) was established in [33].
The discrete maximal -regularity helps us to control the nonlinearlity as well (see [1, 26, 23]), and besides it enables us to obtain optimal-order -norm error estimates without using the inverse inequality (cf. [24, 32, 23]).
Overview of maximal regularity results on surfaces.
The maximal -regularity of parabolic equations on an evolving surface , , as well as the maximal -regularity of time discretizations on an evolving surface and its application to the convergence analysis of BDF methods for nonlinear PDEs on an evolving surface, was discussed in [23].
Semi-discrete maximal regularity results on evolving surfaces.
However, since the maximal -regularity of spatial discretizations for parabolic equations on an evolving surface is still missing, only semi-discretization in time were considered in [23]. The aim of this article is to fill in this gap, by establishing spatially discrete maximal -regularity of isoparametric finite element methods (FEMs) for parabolic equations on an evolving surface which is approximated by quasi-uniform curved triangles.
In order to prove the discrete maximal -regularity for the spatially semi-discrete problems, we combine the techniques developed for evolving surface FEMs and local energy estimates. Firstly, we shall prove the discrete maximal -regularity for spatially semi-discrete FEM in (2.12) on a stationary surface (Theorem 2.2). Then we use a temporal perturbation argument to extend this result to evolving surfaces (Theorem 2.3). The discrete maximal -regularity results for (2.14) can be obtained analogously by a perturbation argument for the lower-order advection term.
Since we are considering a spatially semi-discrete setting, the underlying smooth surface should also be replaced by the finite element surface . The discrepancy between and is the main obstacle in the proof, leading to the following technical difficulties to be addressed:
-
•
The discrete delta functions on and are not simply related by the lift via the distance projection. Indeed, they are correlated via a nonlinear way which stems from the nonlinear relation of projections and . Therefore, it is necessary for us to obtain the high-order consistency between the discrete delta functions on different surfaces (Lemma 3.3) in order to ensure the consistency of the corresponding Green’s functions, which are indispensable in the used local energy estimate (Lemma 3.5) and dyadic decomposition argument (Lemma 4.1).
-
•
The discrepancy of and will introduce a bunch of additional geometric perturbation terms in the local energy estimate (Lemma 3.5) and in the dyadic decomposition argument (Lemma 4.1). We need to treat them carefully to make sure that the leading order stability is still available.
-
•
In the temporal perturbation argument, we need to develop the norm equivalence of the discrete Laplacian (Lemma 5.1 and Remark 5.1). The super-approximation property (cf. (P3) in Section 3.2) also plays an important role in the derivation of this equivalence. Besides, as a nature of parametric finite elements, the matrix-valued coefficient of the following change of variable
|
|
|
always has jumps at the edges and thus is discontinuous. To this end, it is desirable as well to construct a globally continuous substitute according to the definition of the discrete Laplacian.
-
•
The norm equivalence of the discrete Laplacian will bring in a lower-order term . To control this term by the maximal -regularity, we need to use the discrete interpolation inequality (Lemma 5.2) whose proof greatly relies on the -stability of the Ritz projection. The latter is a consequence of the Green’s function estimate on closed manifold ([13, Theorem 3.2]).
The article is organized as follows:
In Section 2, we introduce the basic notations for evolving surface FEMs, the semi-discrete evolving surface FEMs for parabolic equations on an evolving surface, and the main theoretical results about discrete maximal -regularity of evolving surface FEMs.
In Section 3, we develop the preliminary results of geometric perturbation estimates, Green’s function estimates and local energy estimates on the stationary surface. We will prove the maximal regularity on stationary surface (Theorem 2.2) and on evolving surface via a temporal perturbation argument (Theorem 2.3) in Section4 and Section 5, respectively.
In Appendix A we present the detailed proof of local energy estimates (Lemma 3.5), which is a technical lemma used in the proof of the main theorems.
4 Discrete maximal regularity on a stationary surface
(Proof of Theorem 2.2)
We shall prove the following key lemma using the local energy estimates in Section 3.6.
Lemma 4.1
The functions , and
satisfy
|
|
|
(4.1) |
|
|
|
(4.2) |
where the constants and are independent of .
To simplify the notation, in this subsection we continue to omit the dependence of surface on . Additionally, we relax the dependence of the Green’s functions on by denoting
|
|
|
Since the coefficient and are bounded everywhere, the same proofs in [32, Section 4.3–4.4] using Lemma 4.1 can be carried over almost verbatim here.
For example, by differentiating (3.42) in time, we obtain
|
|
|
|
|
|
|
|
|
|
|
|
where and are the linear operators whose kernel are and respectively. Therefore, by using Lemma 4.1 and the arguments in [32, Section 4.3–4.4], we can obtain the following result for :
|
|
|
This proves the maximal regularity result in Theorem 2.2 (by choosing and for any fixed ).
In order to prove Lemma 4.1, we apply the local energy estimate in Lemma 3.5 to estimate . The estimation consists of two parts: The first part concerns estimates for , and the second part concerns estimates for , which is a consequence of the smoothing property of parabolic equations.
Part I.
First, we present estimates in the domain with the restriction ; see (3.45). In this case, the basic energy estimate gives
|
|
|
(4.4) |
|
|
|
(4.5) |
|
|
|
(4.6) |
|
|
|
(4.7) |
|
|
|
(4.8) |
where we have used (3.23) and (3.25) to estimate and , respectively. Hence, we have
|
|
|
(4.9) |
Since the volume of is , we can decompose
in the following way:
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
(4.10) |
where we have used (4.4), (4.6) and (4.7) to estimate
|
|
|
and introduced the following notation:
|
|
|
|
(4.11) |
It remains to estimate . To this end, in view of (3.28) and (3.40),
we set “, , and ”
and “, , and ”
in Lemma 3.5, respectively.
Then we obtain
|
|
|
|
(4.12) |
|
|
|
|
and
|
|
|
|
(4.13) |
|
|
|
|
respectively.
By using interpolation error estimate, (3.33) (exponential decay of ) and Lemma 3.4 (local estimates of regularized Green’s function), we have
|
|
|
|
(4.14) |
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
(4.15) |
|
|
|
|
|
|
|
|
(4.16) |
and
|
|
|
|
(4.17) |
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
(4.18) |
|
|
|
|
|
|
|
|
(4.19) |
By substituting (4.12)-(4) into the expression of in (4.11), we have
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
(4.20) |
Since ,
we can convert the -norm in the inequality above to the -norm:
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
(4.21) |
where we have used and (4.4)-(4.8) to estimate
|
|
|
and used the expression of in (4.11) to bound the terms involving .
Since , we can make sufficiently small by first choosing small enough and then choosing large enough ( can be fixed now and will be determined later). Then the last term on the right-hand side of (4) can be absorbed by the left-hand side. Therefore, we obtain
|
|
|
|
(4.22) |
It remains to estimate . We apply the parabolic duality argument: Let be the solution of the following backward parabolic equation on the domain :
|
|
|
with and .
After testing the above equation by , we get
|
|
|
(4.23) |
where
|
|
|
|
|
|
|
|
|
|
|
|
(4.24) |
Both and can be estimated in the same way as [32, Eq. (5.21), (5.23) and (5.24)]. The only difference is the replacement of flat domain in [32] by surface here. These estimates of and can be written as follows:
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
where in the second-to-last inequality we have used the following pointwise estimate of Green’s function (cf. (3.22)):
|
|
|
|
|
|
|
|
|
where in the first and second inequality we have used the Gagliardo–Nirenberg interpolation inequality with for any integer (cf. [6, Theorem 3.70]) and the local estimates of Green’s function (Lemma 3.4 and Remark 3.3) respectively.
Similarly, by using Lemma 3.3 and (3.22), the argument in [32] leads to the following estimate:
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
We decompose the second and third terms on the right-hand side of (4.23) as follows:
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
(4.25) |
By [32, Eq. (5.26)], we know
|
|
|
(4.26) |
Using the interpolation error estimate,
|
|
|
|
and from Lemma 3.1, Sobolev embedding and Hölder’s inequality, it holds
|
|
|
|
|
|
|
|
According to the estimates above and [32, Eqs. (5.31)–(5.32)] (with therein), we get
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
(4.27) |
Note that (cf. [32, Eq. (5.33)] with therein)
|
|
|
(4.28) |
By substituting (4)–(4.28) into (4.22) we obtain
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
Consequently,
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
By choosing to be large enough ( is determined now), the term will be absorbed by the left-hand side of the inequality above. In this case, the inequality above implies
|
|
|
(4.29) |
Substituting the last inequality into (4) yields
|
|
|
|
(4.30) |
Part II.
For , we first define the finite element space with vanishing average:
and
on which we can define the inverse of the discrete Laplacian operators and respectively.
Lemma 4.2
For any , we define . Then we have the following properties
|
|
|
|
|
|
|
|
Proof.
The first result is a consequence of consistency of and . We define and with and define
|
|
|
Since and , it follows that
|
|
|
(4.31) |
and, similarly
|
|
|
(4.32) |
By definition,
|
|
|
|
|
|
|
|
Applying change of variables and subtraction, we obtain
|
|
|
|
Then we test with and use (4.31)–(4.32), Lemma 3.1, Poincaré inequality and the inverse inequality on to derive
|
|
|
|
|
|
|
|
|
|
|
|
and from the triangle inequality and Poincaré inequality again
|
|
|
|
|
|
|
|
Therefore
, and this proves the first result.
The second result follows from [7, Lemma 4.3]
where we use the elliptic regularity theory and the -stability of Ritz projection on the smooth surface (see Section 3.4).
Denote by the analytical semigroup generated by the linear operator . Then, by the previous lemma with replaced by therein for any fixed , if we define then
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
From the norm equivalence on and , it is straightforward to show the constant of Poincaré inequality (with vanishing average) on is bounded from above by the constant of Poincaré inequality (with vanishing average) on . As a consequence, the smallest eigenvalue of has a lower bound, denoted by , which is uniform w.r.t and only depends on . Hence (the average vanishes due to the differential ),
|
|
|
Similarly, we also have
|
|
|
The proof of Lemma 4.1 is complete.
5 Perturbation arguments for an evolving surface
(Proof of Theorem 2.3)
Following usual notational conventions, we will always identify a finite element function and the corresponding vector collecting its nodal values. Such representation is unique if we have specified the underlying domain. For example, given any , the two integrands of
|
|
|
have the same vector of nodal values, denoted by , but are defined on different domains and . When the underlying domain is specified, is automatically substantialized to a finite element function on that domain. Since all of the quantitative computations in this paper involve either integrals or norms, our notations for finite element functions will always have a unique and clear meaning. For another example, and denote the norms of a finite element function (a nodal vector) on the two different surfaces and , respectively.
Correspondingly, the interpolation operator should be interpreted as the determination of the nodal vector which uniquely corresponds to a finite element function after specifying the underlying surface.
Given any , we first pull back the scheme onto the piecewise polynomial approximate surface and then carry out temporal perturbation argument on .
For the fixed time ,
we define the scalar function and the -valued function with to be the smooth prefactor of the following change of variables via the smooth flow parametrized on , i.e. ,
|
|
|
|
|
|
|
|
Similarly, we define the piecewise smooth scalar function and matrix-valued function with to be the prefactor of the following change of variables via the smooth flow parametrized on , i.e. ,
|
|
|
|
|
|
|
|
and define the piecewise smooth and with to be the prefactor of the following change of variables via the discrete flow parametrized on , i.e. ,
|
|
|
|
|
|
|
|
Since , by the interpolation error estimates, it follows that for any
|
|
|
|
|
|
|
|
and moreover from the geometric perturbation error estimates (cf. [13])
|
|
|
|
|
|
|
|
Hence is a good approximation which is globally continuous and piecewise smooth. This global continuity will allow us to take advantage of the definition of discrete Laplacian. With the definitions above we have, for any ,
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
(5.1) |
where is defined as the Riesz representative of the following linear functional on :
|
|
|
The estimate of is given in the lemma below.
Lemma 5.1
For any , we have the estimate
|
|
|
Proof.
We apply change of variables, Leibniz rule, super-approximation estimate (cf. (P3) in Section 3.2) and the inverse inequality to get
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
where in the second to last inequality we have used the consistency estimate for any , and all of the constants are independent of (possibly depend on ). Thus we conclude the lemma by duality.
Lemma 5.2
The following discrete interpolation inequality holds on the polynomial surface , for all ,
|
|
|
|
Proof.
We define function to be the solution of the following elliptic equation on
|
|
|
or equivalently where is the Ritz projection associated to the finite element space the elliptic operator (see Section 3.4).
Then the elliptic regularity theory says
|
|
|
(5.2) |
Moreover from the inverse inequality
|
|
|
|
|
|
|
|
(5.3) |
Define and we get
|
|
|
|
|
|
|
|
|
|
|
|
and therefore
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
(5.4) |
where in the last line we have used (5) and the elliptic regularity theory .
Hence the norm equivalence, Remark 5.1, (5.2), (5) and the -stability of (see Section 3.4) imply
|
|
|
|
|
|
|
|
|
|
|
|
We complete the proof by absorbing into the left-hand side.
Applying the discrete maximal regularity on the stationary surface with (Theorem 2.19) to (5) and using the norm equivalence between and , we get
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
where in the second inequality we have used Lemma 5.1. Then we apply the norm equivalence, Remark 5.1 and Lemma 5.2 to the inequality above to change the domain from to :
|
|
|
|
|
|
|
|
|
|
|
|
(5.5) |
We define (cf. the proof of Theorem 3.1 and (4.22) in [23])
|
|
|
with
|
|
|
Then (5) implies that
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
where we have used the following estimates in the last line (cf. [23, p. 10])
|
|
|
|
|
|
Thus, by Grönwall’s inequality, it holds that
|
|
|
Therefore, the proof of Theorem 2.3 is complete.
Appendix A Proof of Lemma 3.5
The following lemma is proved in [32, Lemma A.1], which can be easily extended to our current scenario.
Lemma A.1
Suppose that, for any , satisfies
|
|
|
|
|
|
Then we have
|
|
|
|
|
|
|
|
|
where the constant is independent of , and .
It is easy the construct the space-time cut-off function which is constant one on and zero outside . We then define and to be the finite element solution of
|
|
|
(A.1) |
with the initial condition . Obviously, is zero outside as well.
Then,
|
|
|
|
|
|
|
|
|
Testing the above equation with and using Lemma 3.1, temporal integration by parts and Hölder’s inequality (cf. [32, p. 37]), we get
|
|
|
|
|
|
|
|
|
|
|
|
(A.2) |
and, similarly, testing the above equation with leads to
|
|
|
|
|
|
|
|
|
|
|
|
(A.3) |
Applying Young’s inequality to (A) and (A), we obtain
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
Using the triangle inequality and Hölder’s inequality in the temporal direction,
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
(A.4) |
By subtracting (3.52) from (A.1) and using suitably arranged cut-off function, we know that for and
|
|
|
Then we apply Lemma A.1 to and get
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
(A.5) |
where in the last inequality, we use the splitting .
Finally, we complete the proof of Lemma 3.5 by applying the triangle inequality to (A) and (A) and using the fact that both and are zero outside .