1. Introduction
Let
be an arbitrary finite number representing the time,
be a domain to be specified later, and
be a positive number representing the kinematic viscosity. The incompressible Navier-Stokes equations model the dynamic of a viscous and incompressible fluid at constant temperature and with constant density. They are given by the following system of PDE’s posed in
:
The vector field
is the velocity,
is the scalar pressure, and to avoid inessential complications, we set the external force
(but all results presented here can be easily extended to the case of a non vanishing external force, see Remark 4). The first equation is the conservation of linear momentum and the second equation, also called the incompressibility constraint, can be considered as the conservation of the mass, since the density is assumed to be constant. The system (
1) has to be supplemented with initial and boundary conditions. Regarding the initial condition we impose that
with
satisfying the compatibility condition
in
. For the boundary conditions we need to specify the assumptions on the domain. We consider three cases,
,
with
being the three-dimensional flat torus, and
being a bounded domain, whose boundary will be denoted by
; we refer to Assumption 1 for the precise hypotheses on
.
For each of the three different cases we impose the different and natural boundary conditions:
Note that the initial datum will be requested to be tangential to the boundary in the case , and to satisfy the condition and in the other cases. Contrary to the system of compressible Navier-Stokes equations, the pressure p, instead of being obtained through a state equation, is an unknown of the system. This is a consequence of the incompressibility conditions and indeed the pressure can be interpreted as Lagrange multiplier associated with the incompressibility constraint. Note that there are no initial/boundary conditions imposed on the pressure, which (since it appears only as a gradient in the momentum equation) is always determined up to an arbitrary function of time.
Generally speaking, it is very difficult to prove existence and uniqueness of smooth solutions to nonlinear PDE’s. Here, with existence we always mean global in time existence, namely existence on any given time interval
, for arbitrary
. The available theories for weak solutions provide a framework to give a proper meaning to PDE’s, without requiring too much regularity on the solutions and they rely on the theory of generalized functions and distributions. In particular, the landmark idea in the theory of weak solutions is to give up on solving the equations point-wise but trying to solve them in an averaged sense, which is meaningful also from a physical point of view. In the case of fluid mechanics, we expect a very complex behavior by (turbulent) flows appearing in real life, hence we expect to be able to capture only averages of the velocity and pressure, see Reference [
1].
The problem of global existence generally becomes easier since the class of available solution is enlarged and several functional analysis tools can be now used. However, the price to pay to have such a relatively simple existence theory is that the uniqueness problem becomes a very difficult one and many calculation which are obvious when dealing with smooth solutions are not possible or hard to be justified. The three-dimensional incompressible Navier-Stokes is a paradigmatic example of a such situation and the introduction of weak solutions dates back about 100 years ago. In fact, in a series of celebrated papers (the time evolution is treated in Reference [
2]) Jean Leray introduced the notion of weak solution as a mathematical tool, but also with a strong understanding of the physics behind the equations. The theory of weak solutions is also strictly linked with the name of Eberhard Hopf [
3] who gave the first contribution to the problem of existence of weak solutions in a bounded domain, by means of the Faedo-Galerkin method.
It is interesting to observe that many methods and techniques of functional analysis (which are now a common background of graduate students in mathematics) originated from the study of PDE’s and especially from those arising in fluid mechanics. In this note we are trying to explain an extremely limited part of the theory: the existence of (Leray-Hopf) weak solutions. This is a topic at the level of most undergraduate students, with a minimal knowledge of Sobolev spaces and functional analysis (mainly weak convergence and weak compactness), as for instance in the widely used (text)books by Brezis [
4], just to name one. Note also that we try to present a minimal spot in the abstract theory of Navier-Stokes equations, which can be an “appetizer” for students trying to start a serious understanding of (part of) the mathematical fluid mechanics. It is impossible to review what is done on the subject, even only for the mathematical analysis side. Nevertheless many information, at an introductory or more advanced level, can be found in several books, see for instance, just to name a few in alphabetical order [
5,
6,
7,
8,
9,
10,
11].
We think that we will not discourage any reader unfolding the (many) mathematical difficulties of the topic, but –instead– we hope that highlighting the challenges which are typical of mathematical fluid mechanics further interest could be stimulated; To this end we quote the following coming from an interview reported in Reference [
12] in a essay in memory of Jacob Schwartz:
When I asked him [Jacob Schwartz] if there was a subject he had trouble learning, he admitted that there was, namely, fluid dynamics. “It is not a subject that can be expressed in terms of theorems and their proofs,” he said.
Following the above point of view, the first step even in the mathematical analysis of the Navier-Stokes equations is that of giving an appropriate definition of weak solutions, which take into account the functional spaces where it is reasonable find weak solutions, the initial and the boundary conditions. Usually, the functional space to be considered are hinted by the a priori estimate available for the system under consideration. The informal notion of an a priori estimate may be a quantitative bound depending only on the data of the problem, which holds for smooth solutions of the system under consideration, regardless their existence. In particular, for system arising from physics, the a priori estimates usually have a deep physical interpretation.
In the context of the three-dimensional incompressible Navier-Stokes equations the main a priori estimate is indeed the conservation of the energy of the system and is given by the following integral equality:
where the space integral is over the domain under consideration.
The equality (
3) has a very simple
formal proof. Indeed, let
be a smooth solution of (
1) and (
2). By multiplying the momentum equation by
u and integrating over
we get
By integrating by parts and using the divergence free condition and (
2) we get
Then, after integration in time on
with
we get (
3). Note that (
3) gives a quantitative bound depending only on
T, and
of square integrals of the velocity field
u and its gradient
. The energy equality (
3) will serve as motivation for the definition of Leray-Hopf weak solution we will give in
Section 3.
Once a reasonable definition of weak solution is given, to prove global existence one usually exploits what it is know as a
compactness argument, which consists in (1) proving the existence of a sequence of relatively smooth approximating solutions satisfying appropriate uniform estimates; (2) proving that limits of these approximating solutions are effectively weak solution of the problem under consideration. We remark that usually the uniform bounds obtained on the sequence of approximating solutions are the same inferred by the a priori estimates available for the system under consideration; These bounds are then hopefully inherited by weak solutions obtained with a passage to the limit. To be more precise, in the case of the Navier-Stokes equations, the approximation method should be chosen such that the approximate solutions satisfy the energy (in)equality. Due to the limited regularity which can be generally inferred on weak solutions, the validity of any energy balance on the weak solutions to the 3D Navier-Stokes equations is obtained with a limiting process on the approximate solutions and not using the solution
u itself as a test function as done to obtain (
3), since this argument is only formal and not justified when dealing with genuine Leray-Hopf weak solutions.
In this short note we provide a rather self-contained account on the global existence of weak solutions for the three-dimensional incompressible Navier-Stokes equations and some of the (several) approximation methods used in the literature. Since the convergence argument is essentially the same for every approximation methods and for every choice of the domains and boundary conditions mentioned above, we introduce (for the purpose of the exposition) a notion of
approximating solution for which we will prove the convergence to a Leray-Hopf weak solution of the problem (
1) and (
2). This is not the historical path, but is a way we identify to have a unified treatment, which can describe the existence theory within the notion of
approximating solutions.Then, we show how several and well-known approximations fit in the framework introduced and, therefore, we recover the existence of Leray-Hopf weak solution by using those methods. In particular, we will consider the most common techniques available for the construction. Further results based on the energy type methods, concerning uniqueness, regularity and the connection with applied analysis of turbulent flows, can be found in the forthcoming monograph [
1], which is also written in the spirit of being an introduction for undergraduate students, interested in applied analysis of the Navier-Stokes equations.
Organization of the Paper
The paper is organized as follows: In
Section 2 we introduce the functional spaces that we use. Then, in
Section 3 we define of Leray-Hopf weak solutions and study their main properties. In
Section 4 we give the definition of
approximating solution and we prove the convergence to a Leray-Hopf weak solution. Finally, in
Section 5 we prove that certain approximating schemes fit in the framework of
approximating solution.
3. Definition of Leray-Hopf Weak Solutions
In this section we give the definition of Leray-Hopf weak solutions and we prove some related properties. The definition is the following.
Definition 1. A measurable vector field is a Leray-Hopf weak solution of the Navier-Stokes Equations (
1)
and (
2)
if the following conditions are satisfied. - 1.
- 2.
For any and any , it holds that - 3.
Remark 1. It is important to point out that it is an open problem whether or not condition (
7)
can be deduced from the conditions (
5)
and (
6)
. Note also that in the definition we have (
7)
which is the so-called global energy inequality and not the equality (
3)
. Remark 2. In literature Leray-Hopf weak solutions are often defined in the space rather than and satisfying (
7)
for a.e. everywhere instead that for any . This is equivalent to Definition 1, because in that case the velocity field can redefined on a set of measure zero in time in order to lie in and satisfying (
7)
for any , see Reference [14]. We preferred to start with a solution already weakly continuous, to avoid the technical step of redefinition. We want to show that once we have proved the existence of a vector field satisfying the conditions in the Definition 1, we are actually solving the initial value boundary problem (
1) and (
2) in the sense of distributions. First of all we notice that from the condition (
1), we can deduce that
u is divergence-free and satisfies the boundary conditions (
2) in the appropriate weak sense. The following lemma guarantee that
u attains the initial datum
.
Lemma 1. Let and u a Leray-Hopf weak solution. Then, Proof. For
and
, we consider the following function
Then, by using
, after sending
and using that
we arrive to the following estimate:
Then, for any fixed
we can send
and we can conclude that
. By using the Helmholtz decomposition we deduce that this is true for any
and therefore
a.e. on
. Moreover, the previous calculations also show that
and, again by using the Helmholtz decomposition, the same result is valid also for
. By weak lower semi-continuity of norms in weak convergence we get
Next, by using the energy inequality (
7) we also get, by disregarding the non-negative dissipative term and taking the superior limit that
This shows that , which combined with the weak convergence implies the strong convergence, since we are in an Hilbert space. Since the norm induced on H is the same as in , this proves the strong convergence also in H. □
Finally, we show that to any Leray-Hopf weak solution
u it is possible to associate a pressure
p such that
solves the momentum equation in (
1) in the sense of distributions.
Lemma 2. Let u be a Leray-Hopf weak solution of (
1)
and (
2)
. Then, there exists such that and, for any , we have and .
In the case of a general bounded domain
satisfying the Assumption 3, the proof of the Lemma 2 is very technical and requires several preliminaries of operator theory. We refer to References [
7,
10,
11,
15,
16] for the proof. On the other hand in the case of
has no physical boundary the proof is straightforward. We consider here the case
.
Proof. Let
u be a Leray-Hopf weak solution in the sense of Definition 1. For a.e.
consider the elliptic problem
Note that by (
5), Gagliardo-Nirenberg Sobolev inequality (
4), and standard elliptic regularity we can infer that there exists a unique solution of (
8) satisfying
. Next, we show that
solve the Navier-Stokes equations in the sense of distributions. Let
with
and
. Let
be the Helmholtz decomposition, where we denote by
the divergence-free part of
. Then, since
P and
Q commute with derivatives because there are no physical boundaries, we have that
where we have used (
6) in the second equality and (
8) together with the fact that
for some
in the last equality. Finally, by an approximation argument, we have that (
9) holds for any
and we conclude. □
4. Approximate Solutions of the Incompressible Navier-Stokes
Equations
In this section, we define the notion of approximate sequence of solutions to the Navier-Stokes equations and we prove the convergence to Leray-Hopf weak solutions. We use an approach which is a little different from the one usual used. Our choice, which does not follows the historical path, is motivated by the pedagogical purpose of having a unified treatment for several different methods.
Definition 2. Let. We say that is an approximate sequence of solutions with divergence-free initial datumif
- 1.
- 2.
For any and any there exists such that for any - 3.
- 4.
For any and it holds that
Since generally the existence of (smooth) approximating sequences is
rather easy to be proved, the advantage of this definition is that one has just to check a condition on the data and condition (
12) on the remainder (commutator) to show that the approximate solutions converge to a Leray-Hopf weak solution, as is done in the next theorem.
Theorem 1. Let and be a sequence of approximate solutions with initial data such that Then, up to a sub-sequence not relabelled, there exists u such that if Ω satisfies and then Moreover, u is a Leray-Hopf weak solution of (
1) and (
2)
. Remark 3. We stress that because of Remark 1 requiring condition (
13)
(that is a good energy balance already on the approximate functions) is fundamental in order to obtain the energy inequality (
7)
. Moreover, by inspecting the proof below it will be clear that given satisfying (1)
and (2)
in Definition 2 then there exists u satisfying (1)
and (2)
in Definition 1 such that the convergences (
15)
and (
16)
hold. This remark will be important in the analysis of the Implicit Euler Scheme in Section 5.3, because the scheme will not fully fit in the framework of Definition 2. Remark 4. We point out that Theorem 1 is not limited to the case of vanishing external force, but the result holds also in the presence of an external body force . This can be obtained with minor changes in the proof. In this case Definition 2 must also be integrated with an approximating sequence , and adding the term to the left-hand side of (
11)
. This is due to the fact that some of the approximation methods in Section 5 require the body force to be smooth and it is enough to require in . We also note that requiring the body force to be in (and not only in ) is needed in order to remain inside the space of distributions and to have an associated pressure as in Lemma 2; We refer to Reference [17] for more details on this issue. We start with the following straightforward corollary of the classical Arzelà-Ascoli theorem for real functions of a real variable.
Lemma 3. Let E be a separable Banach space and let be a dense subset. Let be a sequence of measurable functions such that . Assume that
- 1.
the sequence is equi-bounded in ,
- 2.
for any fixed the sequence of real functions , , is equi-continuous.
Then, and there exists such that, up to a sub-sequence, A fundamental step in the proof of existence for nonlinear partial differential equations is the proof of certain compactness which allows to get strong convergence in suitable norms. Observe that the a-priori bounds are useful to get weak or weak-* convergences, by means of application of the Riesz representation theorem and –more generally– of Banach-Alaoglu-Bourbaki theorem. On the other hand, since
for a linear operator
T, weak convergence allows to consider linear equations, or more precisely, the linear terms in the equations. On the other hand weak convergence is in general not enough to prove that
Hence, by the a priori estimates we can construct a limit object u, but we still have to show that u is a weak solution of the limiting problem.
To address this point several results have been used. Leray used Helly’s theorem on monotone functions and an ingenious application of Riesz theorem with multiple Cantor diagonal arguments. Hopf used an inequality by Friederichs to handle the Galerkin case. Starting from the work of J.L. Lions [
18] it became common to use the approach by the so-called Aubin-Lions lemma, which is borrowed from the general theory of abstract equations and is based on obtaining some estimates on the time derivative (at least in negative space) of the solution. This latter approach is very flexible, but it requires some non-trivial functional analysis preliminaries to estimate the time-derivative, since instead one can use directly some properties coming from the proper definition of the approximation. Note that in the Definition 1 of weak solution there is no mention to the time-derivative. We will show how to obtain compactness in a elementary way, directly from the weak formulation and thus avoiding the use of time derivatives in Bochner spaces. We believe this may be a simpler approach, at least for presentation to students. We also point out that in certain applications to more complex fluid problems as for instance fluids in a moving domain or non-Newtonian fluids with rheology with time-dependent constitutive law, the proper definition of the time derivative is technically complicated and an approach avoiding the use of this notion becomes particularly welcome.
The next lemma provides a general criterion for strong convergence, which has the advantage to avoid assumptions on the time derivative. Of course, the lemma holds only on bounded domains and therefore we exclude the whole space case, since in the latter one has to work locally. We also stress that the hypothesis are not optimal since we do not prove an if and only if, but the hypotheses are easily verifiable for general nonlinear evolution problems.
The lemma below is very similar to the one proved by Landes and Mustonen in Reference [
19] and for an application to the Navier-Stokes equations see Landes [
20]. For an optimal version (at least in general Hilbert spaces) we refer to Rakotoson and Temam [
21].
Lemma 4. Let be any bounded domain or . Let and assume that and is a sequence such that and weakly in for a.e. . Then, it holds that Proof. We prove the lemma only in the case of being a bounded domain with smooth boundary. First, since and are in , their extensions to zero off U are both in . We denote by and these extensions. a.e. .
Let
be a standard spatial mollifier and set
and
. Next, we have that
and the same estimate holds also for
. Therefore, the following estimates hold
with bounds depending only on
.
Next, by triangular inequality we have
The first term from the right-hand side can be made arbitrarily small by choosing small enough.
To conclude, we first note that by definition of convolution, there exists
such that
Next, since clearly it holds that for the extended functions it holds
we also have that
This follows by fixing
, a time
t such that
,
, and noticing that
as
.
This shows that, for any fixed
, the last term in last inequality in (
17) goes to zero as
, by using Dominated Convergence Theorem. The proof is concluded since we showed that
can be made arbitrarily small. □
The following theorem is the main result of this section.
Proof of the Theorem 1. Let
be a sequence of approximate solutions. By condition (
10) of Definition 2 we can infer that up to a sub-sequence (not relabelled) there exists
such that
For
and
, we consider the following function
Let
with
and
. By using the function
with first with
and then with
, together with the fact that
we can infer that
Next, for
, let
. Then, condition (
10), the Gagliardo-Nirenberg-Sobolev inequality (
4), and the hypothesis on
in (
11) imply that the family
is equi-integrable and then the function
is equi-continuous. Since
is dense in
H and
H is reflexive, we can conclude by using Lemma 3 that
By using (
22) and (
20) we can prove that
u satisfies the energy inequality (
7). Indeed, for any
we have that
Then,
where we have used (
14). In order to conclude it remains only to prove (
15) and (
16). If
is the flat torus or a bounded domain, then (
15) follows directly by Lemma 4. If
we need a localization argument. We first note that by the Helmholtz decomposition (
22) holds also in
. Next, for
let
such that
on
and define
. The sequence
satisfies the hypothesis of Lemma 4, and therefore, after a diagonal argument, it follows that there exists a sub-sequence not relabelled such that
Then, condition (
23) easily implies (
16). □