MnLargeSymbols’164 MnLargeSymbols’171
A stochastic Allen-Cahn-Navier-Stokes system
with singular potential
Abstract.
We investigate a stochastic version of the Allen–Cahn–Navier–Stokes system in a smooth two- or three-dimensional domain with random initial data. The system consists of a Navier–Stokes equation coupled with a convective Allen–Cahn equation, with two independent sources of randomness given by general multiplicative-type Wiener noises. In particular, the Allen–Cahn equation is characterized by a singular potential of logarithmic type as prescribed by the classical thermodynamical derivation of the model. The problem is endowed with a no-slip boundary condition for the (volume averaged) velocity field, as well as a homogeneous Neumann condition for the order parameter. We first prove the existence of analytically weak martingale solutions in two and three spatial dimensions. Then, in two dimensions, we also estabilish pathwise uniqueness and the existence of a unique probabilistically-strong solution. Eventually, by exploiting a suitable generalisation of the classical De Rham theorem to stochastic processes, existence and uniqueness of a pressure is also shown.
Key words and phrases:
Allen-Cahn-Navier-Stokes system; stochastic two-phase flow; logarithmic potential; martingale solutions; pathwise uniqueness.2020 Mathematics Subject Classification:
35Q35, 35R60, 60H15, 76T061. Introduction
Modeling the behavior of immiscible (or partially miscible) binary fluids is a very active area of research because of its importance, for instance, in Biology and Materials Science. A well-known and effective approach is the so-called diffuse interface method (see, e.g., [6]). This approach is based on the introduction of an order parameter (or phase field) which accounts for the presence of the fluid components in a sufficiently smooth way, that is, there is no sharp interface separating them but a sufficiently thin region where there is some mixing. More precisely, denoting by the relative difference between the (rescaled) concentrations of the two components, the regions and represent the pure phases. However, they are separated by diffuse interfaces where can take any intermediate value, i.e. . The interaction between the two components is a competition between the mixing entropy and demixing effects and can be represented by a potential energy density of the form
(1.1) |
for some . This is known as the Flory–Huggins potential (see [37, 55]). Letting be a (sufficiently) smooth domain of , , the Helmholtz free energy associated with the order parameter is then given by
where the first term accounts for the surface energy separating the phases. Here is related to the thickness of the diffuse interface. Then, the functional derivative of is called the chemical potential and usually denoted by , that is,
We can now introduce the two basic equations which govern the evolution of in some time interval : the Cahn-Hilliard equation (see [20, 19])
and the Allen-Cahn equation (see [5])
Here we have assumed that the mobility is constant and equal to the unity. We also recall that, due to the singular behavior of the mixing entropy, the Flory-Huggins potential (1.1) is often approximated with a regular potential like
(1.2) |
This choice simplifies the mathematical treatment. However, when the total mass of is conserved (e.g. in (1.5)) one cannot ensure that takes its values in the physical range . Here we choose to keep the physically relevant potential also in view of extending our analysis to conserved Allen-Cahn equations where in (1.6) or in (1.9) is replaced by , being the spatial average of (see [68], see also [42] and references therein).
When we deal with a two-component fluid mixture, the equation for the phase variable is coupled with an equation for a suitably averaged velocity of the fluid mixture itself. A well known choice is the Navier–Stokes system subject to a capillary force, known as Korteweg force, which can be represented as . More precisely, in the case of an incompressible mixture and taking , constant density equal to the unity and constant viscosity , we have
(1.3) | ||||
(1.4) |
coupled with
(1.5) |
or
(1.6) |
in , for some given . Here represents the volume averaged velocity and stands for the pressure. System (1.3)-(1.4) coupled with (1.5) is known as Cahn–Hilliard–Navier–Stokes system, if (1.5) is replaced by (1.6) then the system is known as Allen–Cahn–Navier–Stokes system. We recall that the standard boundary conditions are no-slip for and no-flux for (1.5) or (1.6).
Starting from the pioneering contribution [53], two-phase flow models have then been developed in several works. In particular, we refer to [48] for the Cahn–Hilliard–Navier–Stokes system and to [14] for the Allen–Cahn–Navier–Stokes system (see [4, 40, 63] for more refined models with unmatched densities and [52, 51] for general thermodynamic derivations). The corresponding mathematical analysis of such models has also experienced a remarkable development in the last decades. Concerning the Cahn–Hilliard–Navier–Stokes system with matched densities see [1, 43] and references therein (see also [2, 3, 12, 41, 44, 45] for more general models). Regarding the Allen–Cahn–Navier–Stokes system, we refer to [38, 39] for the matched case (see also references therein) and to [34, 35, 54, 57, 42, 58, 60, 61] for more refined models.
The deterministic description fails in rendering possible unpredictable oscillations at the microscopic level. These include, for example, the environmental noise due to temperature and configurational effects. The most natural way to take into account such factors was first proposed in [21] where a stochastic version of the Cahn–Hilliard equation was introduced (see also [16, 15] for nucleation effects). That equation has been analyzed under various assssumptions in a number of contributions (see, for instance, [22, 25, 24, 31, 47] and, more recently, [72, 74, 70]). We also refer to [64, 71, 73] for related stochastic optimal control problems. The stochastic Allen-Cahn equation has been investigated in the framework of regular potentials (see, for instance, [13, 50, 49, 65] for examples of well-posedness analysis, see also [18, 69] for numerical schemes and simulations). The singular potential has been analyzed in [10] (see also [9] for the double obstacle potential), while in [11] the separation property from the pure phases has been established.
Here we analyze a stochastic version of the Allen–Cahn–Navier–Stokes system characterized by two independent sources of randomness, the former acting on the fluid velocity and the latter acting on the order parameter dynamic to incorporate thermal fluctuations. More precisely, on account of (1.3)-(1.4) and (1.6), taking for the sake of simplicity, we consider the following system of stochastic partial differential equations
(1.7) | ||||
(1.8) | ||||
(1.9) | ||||
(1.10) | ||||
(1.11) | ||||
(1.12) |
Here and are two independent cylindrical Wiener processes on some (possibly different) separable Hilbert spaces, and is a suitable stochastically integrable process with respect to , for . Moreover, stands for the outward normal unit vector to .
The presence of random terms in both the equations has been considered in [33] in the case of a smooth potential like (1.2) (see also [46, 76, 80] for modified models and [26, 28, 77, 79] for random terms only in the Navier-Stokes system). We also remind that the case of Cahn–Hilliard–Navier–Stokes system for a compressible fluid has been studied in [32] (see, e.g., [78, 27] for random terms only in the Navier-Stokes system in the case of regular potential).
Here, for system (1.7)–(1.12) with a potential like (1.1), we prove the existence of martingale solutions in dimension two and three, as well as pathwise uniqueness and existence of probabilistically-strong solutions in dimension two. The main difficulties on the mathematical side are two. The former is the presence of noise also in the Allen-Cahn equation with singular potential: this requires some ad hoc ideas based on a suitable compensation between the degeneracy of the noise and the blow up of at the endpoints (see ((A3)) below). The latter is the coupling term in the Navier-Stokes equation. Indeed, for the Allen–Cahn equation one can recover only a -regularity for , while for the Cahn–Hilliard equation one gets . This results in the necessity to reformulate the first equation for the fluid in an alternative fashion, i.e., without employing explicitly.
We recall that, in [33], the authors proved the existence of a (dissipative) martingale solution for a similar problem with a smooth potential. Then, taking advantage of the smooth potential, they used the maximum principle to show that the range of the order parameter remains confined in . Thus the global Lipschitz continuity of the potential and its derivatives holds. This fact was exploited to prove the weak-strong pathwise and in law uniqueness in dimension three. However, if the potential is given by (1.1), then no global Lipschitz continuity can be achieved unless one can prove that the solution stays uniformly away from the pure phases, but this is not straightforward in the stochastic case (see [11] for the single stochastic Allen–Cahn equation).
Besides the existence and uniqueness of solutions, there are still a number of issues to investigate, which will be object of future work. For example, regularity properties of the solution and existence of analytically-strong solutions are open issues. The low regularity of the chemical potential in the Allen-Cahn equation that we have mentioned above seems to make the analysis challenging. Yet, some higher regularity properties have been shown in the deterministic case (see [42]). Their extension to the stochastic case is currently under investigation. Moreover, in the spirit of [11], it would be interesting to establish some random strict separation property from the pure phases. To do this, suitable regularity results might be needed. It also worth pointing out that system (1.7)–(1.12) is the non-conserved version of the model, meaning that the spatial average of is not preserved during the evolution. The deterministic conserved version is now well-understood (see [42]). Its stochastic counterpart will also be the subject of further analysis. This issue will require a tuning of the diffusion coefficient (see for example [7]). Finally, we point out that also more general versions of the stochastic Cahn–Hilliard-Navier–Stokes system might be considered on account of the recent advances in the analysis of the stochastic Cahn–Hilliard equation.
The content of this work is structured as follows. In Section 2, we introduce the notation used throughout the work and state the main results. Sections 3 and 4 are devoted to the proof of existence of a martingale solutions and, in dimension two, of a probabilistically-strong solution, respectively.
2. Preliminaries and main results
2.1. Functional setting and notation
For any (real) Banach space , its (topological) dual is denoted by and the duality pairing between and by . If is a Hilbert space, then the scalar product of is denoted by . For every couple of separable Hilbert spaces the space of Hilbert-Schmidt operators from to is denoted by the symbol and endowed with its canonical norm . Let be a filtered probability space satisfying the usual conditions (namely it is saturated and right-continuous), with being a prescribed final time. We will use the symbol to denote identity in law for random variables. Throughout the paper, and are independent cylindrical Wiener process on some separable Hilbert spaces and , respectively. For convenience, we fix once and for all two complete orthonormal systems on and on . We denote by the progressive sigma algebra on . For every and for every Banach space the symbols and indicate the usual spaces of strongly measurable Bochner-integrable functions on and , respectively. For all we write to stress that measurability is intended with respect to . For all and for every separable and reflexive Banach space we also define
which yields by [30, Thm. 8.20.3] the identification
In case of distribution-valued processes, for every , , and we set
Let and consider a bounded domain with smooth boundary and outward normal unit vector . The spatiotemporal domains generated by are denoted by and for all . Moreover, we employ the classical notation , where and , for the real Sobolev spaces and we denote by their canonical norms. We define the Hilbert space , , endowed with its canonical norm , and indicate by the closure of in . We now define the functional spaces
endowed with their standard norms , , and , respectively. As usual, we identify the Hilbert space with its dual through the Riesz isomorphism, so that we have the variational structure
with dense and compact embeddings (both in the cases and ). We will also denote by the variational realization of the with homogeneous Neumann boundary condition, namely
For any Banach space , we use the symbol for the product space . We also need to define the following solenoidal vector-valued spaces
The space is endowed with the norm of and its respective scalar product . By means of the Poincaré inequality, on the space we can use the norm , , induced by the scalar product . The -dimensional realisation of the with homogeneous Dirichlet boundary condition is defined as
Furthermore, we also point out that for any and we have
The Stokes operator is defined as the canonical Riesz isomorphism of , i.e.
Employing the spectral properties of the operator , as customary, we also define the family of operators for any . In particular, if denote the eigencouples of , where is an orthonormal basis of and an orthogonal basis of , we introduce for any the family of Hilbert spaces
and we set . Next, for all , we define the operators
Hereafter, we recall a number of standard facts:
-
(i)
if , then the Hilbert space denotes the so-called part of in ;
-
(ii)
if , then we have and ;
-
(iii)
if , then is the identity operator in so that ;
-
(iv)
if , then coincides with the inverse of the Stokes operator on and extends it on .
In light of the previous considerations, using as pivot space, we also have the general variational structure
for any , with dense and compact embeddings in two and three spatial dimensions. Finally, we remind that, owing to the Korn inequality, we have
where denotes the symmetric gradient. Furthermore, we define the usual Stokes trilinear form on
and the associated bilinear form as
Let us recall that for all
, from which it follows in particular
that for all .
Moreover, we point out that thanks to the usual functional embeddings it holds that
, hence, in particular, that
.
We now report for the reader’s convenience a basic embedding result and its proof. This will be useful in the forthcoming analysis.
Lemma 2.1.
Let and let be a Banach space. For every , there exists such that . In particular, if then is any quantity in , and if then .
Proof.
The embedding holds trivially for every if . The same follows in the case from the chain of embeddings
Let now . If , satisfy
then the fractional Gagliardo-Nirenberg inequality (see [17, Theorem 1]) entails that
for any . Taking into account the embedding
valid for every , we infer that the right hand side of the inequality is finite for every . Moreover, we also get
If we set , then we have
If the claim follows applying the proved inequality to . ∎
Finally, we shall make precise the rigorous interpretation of the stochastic terms (see (1.7) and (1.9)). As a cylindrical process on , , admits the following representation
(2.1) |
where is a family of real and independent Brownian motions. However, it is well known that (2.1) does not converge in , in general. That being said, it always exists a larger Hilbert space , such that with Hilbert-Schmidt embedding , such that we can identify as a -Wiener process on , for some trace-class operator (see [62, Subsections 2.5.1]). Actually, it holds that . In the following, we may implicitly assume this extension by simply saying that is a cylindrical process on . This holds also for stochastic integration with respect to . The symbol
for every progressively measurable process , where is any (real) Hilbert space. The definition is well posed and does not depend on the choice of or (see [62, Subsection 2.5.2]).
2.2. Structural assumptions
The following assumptions are in order throughout the paper.
-
(A1)
The potential is of class with and satisfies
Furthermore, there exists such that
-
(A2)
The operator is linearly bounded in , namely there exists such that
for any . Moreover, taking or , we assume that is -Lipschitz-continuous for some positive constant .
-
(A3)
Setting as the closed unit ball in , the operator satisfies
where the sequence is such that
and
In particular, note that this implies that is -Lipschitz-continuous with respect to the -metric on , and also . With a slight abuse of notation, we will use the symbol
to indicate the operator
Remark 2.2.
Remark 2.3.
If in (A2), then linear boundedness is directly implied by Lipschitz continuity.
2.3. Main results
We first introduce suitable notions of solution for problem (1.7)-(1.11). The first is a martingale solution, the second is a probabilistically-strong solution.
Definition 2.4.
Let and let satisfy
(2.2) | ||||
(2.3) |
A martingale solution to problem (1.7)-(1.11) with respect to the initial datum is a family
where: is a filtered probability space satisfying the usual conditions; are two independent cylindrical Wiener processes on and , respectively; the pair of processes satisfies
(2.4) | ||||
(2.5) | ||||
(2.6) | ||||
(2.7) | ||||
(2.8) |
and, for every and , it holds that
(2.9) | ||||
(2.10) |
Definition 2.5.
Remark 2.6.
The first main result is the existence of a martingale solution.
Theorem 2.7.
Assume (A1)-(A3) and let . Then, for every initial datum satisfying (2.2)-(2.3) there exists a martingale solution to problem (1.7)–(1.12) satisfying the energy inequality
(2.18) |
for every , -almost surely. Here stands for the Lebesgue measure of . Furthermore, there exists such that
(2.19) |
for every , -almost surely. Finally, the following estimate holds:
(2.20) |
Remark 2.8.
The above result still holds if the viscosity depends on in a smooth way and it is bounded from below by a positive constant. Moreover, we recall that, in [33], the energy inequality is written in a distributional sense.
The second is a stronger result in dimension two, namely, the existence of a (unique) probabilistically-strong solution.
Theorem 2.9.
Assume (A1)-(A3), let , , and in (A2). Then, for every initial datum satisfying (2.2)–(2.3), there exists a unique probabilistically-strong solution for problem (1.7)–(1.12) and a pressure , which satisfy on the original probability space the analogous of the energy inequality (2.18), the pressure-variational formulation (2.19), and the estimate (2.20).
3. Proof of Theorem 2.7
Here we prove the existence of martingale solutions to problem (1.7)–(1.12). For the sake of clarity, the proof is split into several steps.
3.1. Regularization of the singular potential
First of all, note that assumption (A1) implies that the function
can be identified with a maximal monotone graph in . Consequently, one can consider, for every , the resolvent operator and the Yosida approximation of , defined as follows
For notation and general properties of monotone operators we refer the reader to [8]. For every , we define an approximation of as follows
(3.1) |
Thus it holds
(3.2) |
In order to preserve the scaling of the Yosida-approximation on , we analogously define the -approximation of by setting
(3.3) |
Notice that, by assumption (A3) and the non-expansivity of , the operator is -Lipschitz-continuous (therefore uniformly in ), and converges pointwise to as . Now, we consider the -approximated (formal) problem
(3.4) | ||||
(3.5) | ||||
(3.6) | ||||
(3.7) | ||||
(3.8) | ||||
(3.9) |
3.2. Faedo-Galerkin approximation
A discretization scheme is now applied to problem (3.4)-(3.9). Let us consider the (countably many) eigencouples of the negative Laplace operator with homogeneous Neumann boundary condition, namely the couples such that
Analogously, we also consider the (countably many) eigencouples of the Stokes operator, namely the couples , and such that
It is well known that, up to a renormalization, the set (resp. ) is an orthonormal system in (resp. ) and an orthogonal system in (resp. ). Let and consider the finite-dimensional spaces and , both endowed with the -norm. In order to approximate the stochastic perturbation, we define the operators and as
and such that
for any , and . Notice that, fixed any and , and are actually well defined as elements of and , respectively. Indeed, for instance,
(3.10) |
Moreover, since is -Lipschitz continuous in the sense of assumption (A2) and the orthogonal projection on is non-expansive as an operator from to itself, we can deduce by the same argument that is also -Lipschitz continuous as an operator from to . Similar considerations also apply to . More precisely, we have
Proposition 3.1.
Let and . The operators
are well defined and uniformly Lipschitz continuous with respect to and . In particular, is -Lipschitz continuous from to and is -Lipschitz continuous from to .
Next, we define suitable projections (orthogonal with respect to the -inner products) of initial data (evaluated at some point in ) on the discrete spaces and , namely, for all , we set
It is now possible to formulate the discretized problem, which reads
(3.11) | ||||
(3.12) | ||||
(3.13) | ||||
(3.14) | ||||
(3.15) | ||||
(3.16) |
The variational formulation of problem (3.11)-(3.16) is given by
(3.17) | ||||
(3.18) |
for every and . Fixed any and , we search for a weak solution to (3.17)-(3.18) of the form
(3.19) |
where
are suitable stochastic processes. Inserting (3.19) into (3.17)-(3.18) and choosing as test functions and for each , we deduce that the three processes , and satisfy the system of ordinary stochastic differential equations
(3.20) | |||
(3.21) | |||
(3.22) | |||
(3.23) | |||
(3.24) |
Let us point out that, in order to derive (3.20)-(3.24), we exploited the fact that, for every choice of integers and between and ,
(3.25) |
as well as the orthogonality in of the basis . The stochastic integrals in (3.20)-(3.21) have to be regarded as and for every , where
and
for every . By Lipschitz continuity of all the nonlinearities appearing in (3.20)-(3.24), the standard theory of abstract stochastic evolution equations applies. Therefore, we are able to infer that
3.3. Uniform estimates with respect to
First of all, we prove some uniform estimates with respect to the Galerkin parameter , keeping fixed. Hereafter, the symbol (possibly numbered) denote positive constants whose special dependencies are explicitly pointed out when necessary. ‘In some cases, in order to ease notation, we may use the same symbol to denote different constants throughout the same argument. In any case, such constants are always independent of .
First estimate
We exploit the Itô formula for the -norm of given in [62, Theorem 4.2.5]. This gives
(3.26) |
Let us now address the above equality term by term. First of all, recalling (3.2) and that , we find
(3.27) |
Next, owing to (3.10) and (A3), we have
(3.28) |
Finally, by 1-Lipschitz-continuity of the projection , it follows
(3.29) |
Thus, combining (3.27)-(3.29) with (3.26), letting , multiplying the resulting inequality by two, taking -powers, the supremum on the interval and expectations, we arrive at
where depends on and also on , , , . The Burkholder-Davis-Gundy and Hölder inequalities jointly with (3.28) entail
(3.30) |
where only depends on . In turn, thanks to (3.30) and the Young inequality, we can refine the estimate and get
The Gronwall lemma entails that there exists , independent of and , but depending on and the structural data of the problem, such that
(3.31) |
for every fixed .
Second estimate.
We devise a similar argument for the -norm of . Still exploiting [62, Theorem 4.2.5], the Itô formula implies
(3.32) |
Next, we want to apply the standard Itô formula to the regularized energy functional
However, notice that exactly contains the kinetic energy contribution linked to the fluid velocity field which we just handled in (3.32). Thus, it is sufficient to apply the Itô formula only to the portion of the energy linked to the order parameter . Let us stress that this is only possible since no coupling energy terms are present. We set
It has already been shown in [70, Subsection 3.2] that is twice Fréchet-differentiable. Thus it is possible to apply the Itô formula in its classical version [23, Theorem 4.32]. This yields
(3.33) |
where we recall that . Adding (3.32) and (3.33) together, we find
(3.34) |
Fix now . Taking -powers, supremum over , and expectations of both sides of (3.34) yield
(3.35) |
where only depends on . Next, we address the terms on the right hand side of (3.35). By (3.10) and Assumption (A2), we deduce
(3.36) |
Since , recalling assumption (A3), (3.28), and the non-expansivity of , we have
(3.37) |
Furthermore, since , by (3.2) we have that, for all ,
Thus, thanks to (A3) and the non-expansivity of , we get
(3.38) |
Finally, we address the stochastic integrals. Using (3.36) jointly with the Burkholder-Davis-Gundy and Young inequalities, for every we obtain
(3.39) |
where only depends on , , and . Moreover, by (3.28) and the same inequalities, we also get
(3.40) |
where only depends on , , and . Finally, the non-expansivity of the orthogonal projectors on and imply
(3.41) |
whereas, since is linearly bounded, being Lipschitz-continuous, is quadratically bounded so that
(3.42) |
where depends on and . Collecting (3.36)-(3.42) and choosing small enough, from (3.35) we infer that
(3.43) |
Here depends on and . An application of the Gronwall lemma entails the existence of , depending on , and , such that
(3.44) | |||
(3.45) | |||
(3.46) |
Further estimates.
The Lipschitz-continuity of and the fact that entail
for some only depending on . Therefore, thanks to (3.31) we also get the estimate
(3.47) |
Additionally, by comparison in (3.14), we get
(3.48) |
Here or depend on , , and . In light of (3.36), (3.37) and on account of (3.44) and (3.45), we deduce
(3.49) | ||||
(3.50) |
Here, again, the constants , depend on . As a consequence of [36, Lemma 2.1], the following estimates on the Itô integrals hold:
(3.51) | ||||
(3.52) |
for every and , where and depend on , and . Estimates (3.51) and (3.52) enable us to carry out two comparison arguments. Let us interpret (3.18) as an equality in ,
for all such that , , -almost-surely. It is clear that, by the Hölder inequality,
(3.53) |
implying (see (3.44), (3.46), and (3.48))
(3.54) |
for some depending on and . Then, recalling that
and estimate (3.52) as well as Lemma 2.1, we find
(3.55) |
for some if , and for all if . The constant may depend on , and .
Remark 3.3.
Observe that is always well defined. Here, we apply Lemma 2.1 with and . If denotes the Sobolev fractional exponent given by Lemma 2.1, then the following alternative holds:
-
(a)
if , then any value of is valid, and therefore we can set an arbitrary ;
-
(b)
if , then any value of is valid, and therefore we can set an arbitrary .
Similarly, we consider the weak formulation of the discretized Navier–Stokes equation
for all such that , , -almost-surely. Then, we have
Owing to (3.44) and the continuity of , we infer
for some depending on and , but independent of . Next, we recall the well-known inequality
(3.56) |
Therefore, we find
Furthermore, since by the Hölder, Gagliardo–Nirenberg and Young inequalities, we have
(3.57) |
for both and . Thus we get
Summing up, also owing to (3.50) and Lemma 2.1, we arrive at
(3.58) |
for some if , and for all if . Here depends on , and .
Remark 3.4.
In the following, we assume that, given , the exponents and are fixed. Notice that if , then trivially and are both greater than 1.
3.4. Passage to the limit as
Owing to the previously proven uniform estimates, we now pass to the limit as keeping fixed. Let .
Lemma 3.5.
The family of laws of is tight in the space for any . The family of laws of is tight in the space .
Proof.
To prove the claims, we follow a standard argument (refer, for instance, to [70, Subsection 3.3] or [80, Proposition 1]). We first recall we have that the embeddings (see [75, Corollary 5])
are compact (the intersection spaces are endowed with their canonical norm), since , . Here, . Let us prove the first one, the other three cases being similar. For any , let denote the closed ball of radius in . Then the Markov inequality, jointly with estimates (3.45) and (3.58), implies
for some depending on > This yields
so that the first claim is proven. The remaining claims can be proven analogously, replacing the spaces accordingly and exploiting the corresponding estimates. ∎
Lemma 3.6.
The family of laws of is tight in the space . The family of laws of is tight in the space .
Proof.
Next, we consider the constant sequences of cylindrical Wiener processes
Lemma 3.7.
The family of laws of is tight in . The family of laws of is tight in .
Proof.
It directly follows from the fact that every measure on a complete separable metric space is tight. ∎
Finally, we consider the sequences of approximated initial conditions.
Lemma 3.8.
The family of laws of is tight in . The family of laws of is tight in .
Proof.
It is a third iteration of the proof of Lemma 3.5, exploting the compact embeddings
and the Markov inequality on closed balls of and , respectively. ∎
As an immediate consequence of Lemmas 3.5-3.8, we get that the family of laws of
is tight in the product space
Owing to the Prokhorov and Skorokhod theorems (see [56, Theorem 2.7] and [81, Theorem 1.10.4, Addendum 1.10.5]), there exists a probability space and a sequence of random variables such that the law of is for every , namely (so that composition with preserves laws), and the following convergences hold
for some limiting processes belonging to the specified spaces. Let us recall that, for the sake of what follows, if is a finite positive measure space and is any Banach space, then the Bochner space is reflexive if and only if and are reflexive (see, for instance, [29, Corollary 2, p. 100]). By the previously proven uniform estimates and the preservation of laws under , up to a subsequence which we do not relabel, the Vitali convergence theorem, the Eberlein-Smulian theorem and the Banach-Alaoglu theorem entail
Let us now define
By uniform boundedness and weak compactness, there exists some such that
Let us notice that it is possible to take the probability space large enough so that it does not depend on . Taking into account the previous considerations and further straightforward weak convergences, the limit processes fulfill the following regularity properties:
From this starting point, we now address several issues.
The nonlinearities
First of all, by Lipschitz-continuity of , it follows that
Moreover, since is uniformly Lipschitz-continuous (recall Proposition 3.1) and
we conclude
A very similar computation also shows
Next, we address the Korteweg term representing the capillary force. Let us prove that
Indeed, for any ,
and both terms tend to zero as by the above convergences (note that ). Here, stands for the expectation with respect to the probability . As far as the other nonlinear term appearing in the Navier-Stokes equations, we have, as a straightforward application of (3.56),
Finally, we address the convective term. Observe that
Thus, by the Hölder inequality, it holds that
The stochastic integrals
Let us now identify and . The procedure is standard, for instance see [23, Section 8.4]. We introduce a family of filtrations on , namely we set
for any , and , in such a way that both and are adapted. In particular, by preservation of laws and the definitions of Wiener process and stochastic integral, we readily have that is a -Wiener process on and
are respectively a -valued and an -valued martingale. Let us iterate the same procedure on the limit processes: we define
It is easy to infer, by the proven convergences, that both and are zero. Let now , and set
Let be a bounded and continuous function. By definition of martingale, we have
(3.59) |
for . Here, the arguments of are intended to be restricted over when necessary and denotes the expectation with respect to . Letting in (3.59), an application of the dominated convergence theorem, owing to the proven convergences and the properties of , entails
(3.60) |
which expresses the fact that is a -valued -martingale for . The characterization of -Wiener processes given in [23, Theorem 4.6] leads us to compute the quadratic variation of . To this end, notice that (3.60) means that, for every
and using once more the dominated convergence theorem, we get
namely
which is enough to conclude that is a -Wiener process, adapted to , owing to [23, Theorem 4.6]. We are now in a position to study the stochastic integrals. Arguing exactly as in (3.59)-(3.60), we find that (resp. ) is a -valued (resp. an -valued) martingale. As far as the quadratic variations are concerned, an application of [23, Theorem 4.27] yields
for every . Let us outline the argument for the first sequence (similar considerations hold for the second one). Once again, fixing , we have
and, as , the dominated convergence theorem implies that
Notice that in the above equality the dualities are necessary. The quadratic variation of is therefore
Let us identify with the martingale
which is a -valued -martingale having the same quadratic variation of . By [66, Theorem 3.2], we can write
(3.61) |
Thus, we now compute the cross quadratic variation appearing on the right hand side in (3.61). To this end, notice that by [66, Theorem 3.2], we have
where we also used the fact that , where is the classical Hilbert-Schmidt embedding. This implies that
A further application of the dominated convergence theorem entails that, as ,
(3.62) |
Identification of the limit solution.
We are now left to prove that the limiting processes solve the regularized Allen-Cahn-Navier-Stokes system (3.4)-(3.9). Testing (3.11) by some and integrating the obtained identity with respect to time yield
Letting , owing to above convergences and using the dominated convergence theorem, we obtain
(3.63) |
Next, we identify the limit chemical potential. Testing (3.14) by some , passing to the limit as yields and exploiting the proven convergences entail
(3.64) |
almost everywhere in and -almost surely. Finally, consider the approximating Allen-Cahn equation. Testing (3.13) by some and passing to the limit as , we get
Therefore, system (3.4)-(3.9) is satisfied (in the weak sense) once we identify (the law of) the initial state. By the properties of , we know that
for any , and by uniqueness of the distributional limit (jointly with the above convergences) we conclude
The initial conditions are therefore attained in law.
3.5. Uniform estimates with respect to
Here, we prove further uniform estimates, now independent of the Yosida parameter . The symbol (possibly numbered) denotes a positive constant, always independent of , which may change from line to line.
First estimate.
Notice that the constant in (3.31) does not depend on . By lower semicontinuity and preservation of laws of , we infer
(3.65) |
Second estimate.
Let us collect, in (3.35), all controls which are already uniform with respect to , that is, the bounds on the diffusion coefficients (3.36) and (3.37), the bounds on the stochastic terms (3.39) and (3.40), and the initial data bounds given in (3.41). This can be summarized as follows (we can express the result in the new variables since preserves laws)
(3.66) |
where depends on but is independent of . Next, we would like to take the limit as in (3.66). On the left hand side, the previously proven uniform estimates, convergences and weak lower semicontinuity of the norms are enough to pass to the limit. Moreover, it is easily seen, by Lipschitz-continuity of , that in by the dominated convergence theorem. Finally, in order to pass to the limit in the last term at right hand side, we bound each term of the sequence as follows:
Thanks to the proven convergences, it is straightforward to conclude that (cfr. [70])
almost everywhere in . Therefore, applying the dominated convergence theorem and the weak lower semicontinuity of the norms, we find
(3.67) |
We now need to find uniform bounds with respect to for the two terms involving . Notice first that, as customary,
which is finite by the hypotheses on the initial datum. Concerning the other term, we have
where we made use of (A3) and we exploited the non-expansivity of . Collecting the two results in (3.67), we get
(3.68) |
and an application of the Gronwall lemma to (3.68) gives
(3.69) | |||
(3.70) | |||
(3.71) |
Further estimates.
Choosing in (3.64) yields:
and exploiting the monotonicity of , the Hölder and the Young inequalities, after an integration over , we get
Therefore, by estimates (3.65) and (3.71), we find
(3.72) |
Again, by comparison in (3.7), we also obtain
(3.73) |
The remaining estimates can be obtained following line by line the work already showed in Subsection 3.3. In this way, we also recover the following: given any and , there exist and , satisfying and if (see Remarks 3.3 and 3.4), such that
(3.74) | ||||
(3.75) | ||||
(3.76) | ||||
(3.77) |
3.6. Passage to the limit as
We are now in a position to let (along a suitable subsequence). The argument is similar to the one of Subsection 3.4, thus we will omit some details for the sake of brevity. Iterating the proofs of Lemmas 3.5-3.7, we learn that the family of laws of
is again tight in the product space
Here, we recall that and we set and for and any . Owing to the Prokhorov and Skorokhod theorems (see [56, Theorem 2.7] and [81, Theorem 1.10.4, Addendum 1.10.5]), there exists a probability space and a family of random variables such that the law of is for every , namely (so that composition with preserves laws), and the following convergences hold as :
for some limiting processes satisfying
Again, by estimate (3.71), we also have the following weak convergence of the redefined chemical potentials
Mimicking the arguments illustrated in Subsection 3.4, we now address several issues.
The nonlinearities.
First of all, we show that
This comes from the weak-strong closure of maximal monotone operators (see, for instance, [8, Proposition 2.1]) combined with the strong convergence for proved above (recall that . Next, the diffusion coefficients. As for , it is easy by Lipschitz continuity to deduce
Moreover, arguing similarly (recall also Proposition 3.1), we get
and we conclude
Regarding the convective term and the Korteweg force, on account of the obtained convergences, we deduce that
The stochastic integrals.
Following line by line the argument presented in Subsection 3.4, it is possible to identify the limits and . Indeed, we have
which are a and an -valued martingale, respectively, adapted with respect to a suitable filtration .
Identification of the limit solution.
Again, a multiple application of the dominated convergence theorem allows us to infer that the limit processes form a martingale solution of the original problem. The existence of a martingale solution is proved.
3.7. The energy inequality
We are left to prove the energy inequality. To this end, we simply pass to the limit in a suitable approximating energy inequality. Let us add (3.32) and (3.33) together and take expectations. Recalling that stochastic integrals are martingales, we obtain the identity
(3.78) |
Thank to (3.36) and (3.37), from (3.78) we infer
(3.79) |
Exploiting the preservation of laws by , and letting , we find
(3.80) |
Here we have used the lower semicontinuity of the norms and the dominated convergence theorem. A second passage to the limit entails the claimed inequality. Indeed, exploiting preservation of laws by in (3.80) as well as (A3), and letting , we get
(3.81) |
Observe that, passing in the limit in the third term on the left hand side of (3.80) is possible by lower semicontinuity since recalling that
it follows almost everywhere in . Fixed any , the energy inequality follows taking the supremum over in both sides of (3.81).
3.8. Recovery of the pressure.
It is possible to recover a pressure through a generalization of the classical De Rham theorem to stochastic processes (see [59]). The result is of independent interest and we report it hereafter for reader’s convenience.
Theorem 3.9 ([59, Theorem 4.1]).
Let be a bounded Lipschitz domain of and let be a complete probability space. Let and . Let
be such that, for all satisfying ,
Then there exists a unique (up to a constant)
such that
and
Furthermore, there exists a positive constant , independent of , such that
Let us now find suitable values for the parameters and . By choosing with in (1.7), after elementary rearrangements and integration by parts we obtain that
for almost every , -almost surely. Hence, by setting
one has in particular, for all with , that
Let us recover the regularity of . Observing that and that is linear and continuous, we have
Furthermore, recalling that thanks to the fundamental theorem of calculus as shown in the proof of [59, Theorem 2.2], one has that
Moreover, since for the bilinear form
is continuous, thanks to the regularity of it follows that
Eventually, iterating the computations in (3.57), we obtain
Hence, we have shown that and an application of Theorem 3.9 with , and yields the existence of the (unique up to a constant) pressure . Finally, we derive an estimate for . The continuous dependence given by Theorem 3.9 implies that
Knowing that
and exploiting the Burkholder-Davis-Gundy inequality together with assumption (A2), we arrive at
The proof of Theorem 2.7 is complete.
4. Existence of probabilistically-strong solutions when
This section is devoted to proving Theorem 2.9. To this end, we will use a standard approach, namely we shall deduce it from pathwise uniqueness of martingale solutions.
Proposition 4.1.
Let and . Assume (A1)-(A3) and consider two sets of initial conditions for complying with the hypotheses of Theorem 2.7. Let denote some martingale solutions to (1.7)-(1.12), defined on the same suitable filtered space and with respect to a pair of Wiener processes . Then, there exist a sequence of positive real numbers and a sequence of stopping times , with -almost surely as , such that the following continuous dependence estimate holds
In particular, the martingale solution to (1.7)-(1.12) is pathwise unique.
Proof.
Let us set
For every and we define the stopping time as
with the usual convention that , and set
Clearly, almost surely as . Let us also introduce the functionals
We point out, once and for all, that what follows is valid -almost surely for every . Let us consider at first . First of all, let us compute its first two Fréchet derivatives. If we set
then we have . Therefore, an application of the chain rule implies that is defined by
Here, of course, we exploited the facts that and that . The above identity must be understood as follows
Moreover, by the properties of the inverse of the Stokes operator, it holds
(4.1) |
for every . Notice that and thus for every . Applying the Itô lemma [23, Theorem 4.32] to and stopping at time , we obtain
(4.2) |
For the ease of notation, throughout computations we may omit the evaluation of the functions at the time , for -almost every . We address the various terms in (4.2) separately. First of all, notice that, by (4.1),
on account of the incompressibility condition
for . Then, using the Hölder, Young and Ladyzhenskaya inequalities, together with the definition of , we find
(4.3) |
Here, we also used the well-known fact that is an equivalent norm in . Next, we address the coupling term. We make use of the customary formula
for . The above makes sense in , since the chemical potential is not regular enough. Therefore, integrating by parts, we recover the identities
On the other hand, by Hölder, Young and Ladyzhenskaya inequalities, we obtain
(4.4) |
By Assumption (A2) we also get (recall that ),
(4.5) |
since is an equivalent norm in . Collecting (4.3)-(4.5), we infer from (4.2) that
(4.6) |
Before dealing with the stochastic integral in (4.6), we consider . Applying the Itô lemma to yields, thanks to [62, Theorem 4.2.5],
(4.7) |
Observe now that, by the mean value theorem and (A1),
(4.8) |
Moreover, we have
(4.9) |
By (A3), we easily deduce
(4.10) |
On account of (4.8)-(4.10), from (4.7) we arrive at
(4.11) |
Adding (4.6) and (4.11) together, we obtain
(4.12) |
so that the Gronwall Lemma and the definition of yield
(4.13) |
Take now -powers, the supremum (with respect to time) and expectations (with respect to ): let us deal with the stochastic integrals on the right hand side of (4.13). The Burkholder-Davis-Gundy inequality combined with the Young inequality and (A2) entail, for every , that
(4.14) |
while the same inequalities and (A3) also yield
(4.15) |
Taking (4.14) and (4.15) into account in (4.13) and choosing small enough, an application of the Gronwall lemma entails the claimed continuous dependence estimate. In turn, upon choosing and , this also yields and on the stochastic interval for every . Hence pathwise uniqueness of the solution follows since almost surely. ∎
The existence of a probabilistically-strong solution follows from standard results (see, for instance, [67, Theorem 2.1]), which also turns out to be unique. The existence and uniqueness (up to a constant) of a pressure can be deduced arguing as in Subsection 3.8. The proof of Theorem 2.9 is finished.
Acknowledgments. The second and third authors are members of Gruppo Nazionale per l’Analisi Matematica, la Probabilità e le loro Applicazioni (GNAMPA), Istituto Nazionale di Alta Matematica (INdAM). The second author has been partially funded by MIUR-PRIN Grant 2020F3NCPX “Mathematics for industry 4.0 (Math4I4)”.
References
- [1] H. Abels “On a diffuse interface model for two-phase flows of viscous, incompressible fluids with matched densities” In Arch. Ration. Mech. Anal. 194.2, 2009, pp. 463–506 DOI: 10.1007/s00205-008-0160-2
- [2] H. Abels, D. Depner and H. Garcke “Existence of weak solutions for a diffuse interface model for two-phase flows of incompressible fluids with different densities” In J. Math. Fluid Mech. 15, 2013, pp. 453–480
- [3] H. Abels and E. Feireisl “On a diffuse interface model for a two-phase flow of compressible viscous fluids” In Indiana Univ. Math. J. 57.2, 2008, pp. 659–698 DOI: 10.1512/iumj.2008.57.3391
- [4] H. Abels, H. Garcke and G. Grün “Thermodynamically consistent, frame indifferent diffuse interface models for incompressible two-phase flows with different densities” In Math. Models Methods Appl. Sci. 22, 2012, pp. 1150013
- [5] S. M. Allen and J. W. Cahn “A microscopic theory for antiphase boundary motion and its application to antiphase domain coarsening” In Acta Metall. 27, 1979, pp. 1085–1095
- [6] D.-M. Anderson, G.-B. McFadden and A.-A. Wheeler “Diffuse-interface methods in fluid mechanics” In Annu. Rev. Fluid Mech. 30, 1997, pp. 139–165
- [7] D. C. Antonopoulou, P. W. Bates, D. Blömker and G. D. Karali “Motion of a Droplet for the Stochastic Mass-Conserving Allen–Cahn Equation” In SIAM J. Math. Anal. 48.1, 2016, pp. 670–708
- [8] Viorel Barbu “Nonlinear differential equations of monotone types in Banach spaces”, Springer Monographs in Mathematics Springer, New York, 2010, pp. x+272 DOI: 10.1007/978-1-4419-5542-5
- [9] C. Bauzet, E. Bonetti, G. Bonfanti, F. Lebon and G. Vallet “A global existence and uniqueness result for a stochastic Allen-Cahn equation with constraint” In Math. Methods Appl. Sci. 40.14, 2017, pp. 5241–5261 DOI: 10.1002/mma.4383
- [10] F. Bertacco “Stochastic Allen–Cahn equation with logarithmic potential” In Nonlinear Anal. 202, 2021, pp. 112122
- [11] F. Bertacco, C. Orrieri and L. Scarpa “Random separation property for stochastic Allen-Cahn-type equations” In ArXiv e-prints, 2021 arXiv:2110.06544 [math.PR]
- [12] A. Berti, V. Berti and D. Grandi “Well-posedness of an isothermal diffusive model for binary mixtures of incompressible fluids” In Nonlinearity 24.11, 2011, pp. 3143–3164 DOI: 10.1088/0951-7715/24/11/008
- [13] L. Bertini, P. Buttà and A. Pisante “Stochastic Allen–Cahn equation with mobility” In Nonlinear Differ. Equ. Appl. 24.54, 2017
- [14] T. Blesgen “A generalization of the Navier-Stokes equations to two-phase flows” In Journal of Physics D: Applied Physics 32.10 IOP Publishing, 1999, pp. 1119–1123 DOI: 10.1088/0022-3727/32/10/307
- [15] D. Blömker, B. Gawron and T. Wanner “Nucleation in the one-dimensional stochastic Cahn-Hilliard model” In Discrete Contin. Dyn. Syst. 27.1, 2010, pp. 25–52
- [16] D. Blömker, S. Maier-Paape and T. Wanner “Second Phase Spinodal Decomposition for the Cahn-Hilliard-Cook Equation” In Trans. Amer. Math. Soc. 360.1, 2008, pp. 449–489
- [17] H. Brezis and P. Mironescu “Gagliardo–Nirenberg inequalities and non-inequalities: The full story” In Ann. Inst. H. Poincaré Anal. Non Linéaire 35.5, 2018, pp. 1355–1376
- [18] C.-B. Bréhier and L. Goudenège “Analysis of some splitting schemes for the stochastic Allen-Cahn equation” In Discr. Cont. Dyn. Syst. Ser. B 24.8, 2019, pp. 4169–4190
- [19] J. W. Cahn and J. E. Hilliard “On spinodal decomposition” In Acta Metall. 9, 1961, pp. 795–801
- [20] John W. Cahn and John E. Hilliard “Free Energy of a Nonuniform System. I. Interfacial Free Energy” In The Journal of Chemical Physics 28.2, 1958, pp. 258–267 DOI: 10.1063/1.1744102
- [21] H.E Cook “Brownian motion in spinodal decomposition” In Acta Metallurgica 18.3, 1970, pp. 297 –306 DOI: https://doi.org/10.1016/0001-6160(70)90144-6
- [22] Giuseppe Da Prato and Arnaud Debussche “Stochastic Cahn-Hilliard equation” In Nonlinear Anal. 26.2, 1996, pp. 241–263 DOI: 10.1016/0362-546X(94)00277-O
- [23] Giuseppe Da Prato and Jerzy Zabczyk “Stochastic equations in infinite dimensions” 152, Encyclopedia of Mathematics and its Applications Cambridge University Press, Cambridge, 2014, pp. xviii+493 DOI: 10.1017/CBO9781107295513
- [24] Arnaud Debussche and L. Goudenège “Stochastic Cahn-Hilliard equation with double singular nonlinearities and two reflections” In SIAM J. Math. Anal. 43.3, 2011, pp. 1473–1494 DOI: 10.1137/090769636
- [25] Arnaud Debussche and Lorenzo Zambotti “Conservative stochastic Cahn-Hilliard equation with reflection” In Ann. Probab. 35.5, 2007, pp. 1706–1739 DOI: 10.1214/009117906000000773
- [26] G. Deugoué, B. Jidjou Moghomye and T. Tachim Medjo “Approximation of a stochastic two-phase flow model by a splitting-up method” In Commun. Pure Appl. Anal. 20, 2021, pp. 1135–1170
- [27] G. Deugoué, A. Ndongmo Ngana and T. Tachim Medjo “Strong solutions for the stochastic Cahn-Hilliard-Navier-Stokes system” In J. Differential Equations 275, 2021, pp. 27–76
- [28] G. Deugoué and T. Tachim Medjo “Large Deviation for a 2D Allen–Cahn–Navier–Stokes Model Under Random Influences” In Asymptot. Anal. 123.1-2, 2021, pp. 41–78
- [29] J. Diestel and J.J. Uhl “Vector Measures”, Mathematical surveys and monographs American Mathematical Society, Providence, RI, 1977
- [30] R. E. Edwards “Functional analysis. Theory and applications” Holt, RinehartWinston, New York-Toronto-London, 1965, pp. xiii+781
- [31] N. Elezović and A. Mikelić “On the stochastic Cahn-Hilliard equation” In Nonlinear Anal. 16.12, 1991, pp. 1169–1200 DOI: 10.1016/0362-546X(91)90204-E
- [32] E. Feireisl and Madalina Petcu “A diffuse interface model of a two-phase flow with thermal fluctuations” In Appl. Math. Optim. 83.1, 2021, pp. 531–563 DOI: 10.1007/s00245-019-09557-2
- [33] E. Feireisl and Madalina Petcu “Stability of strong solutions for a model of incompressible two-phase flow under thermal fluctuations” In J. Differential Equations 267.3, 2019, pp. 1836–1858 DOI: 10.1016/j.jde.2019.03.006
- [34] E. Feireisl, Madalina Petcu and Dalibor Pražák “Relative energy approach to a diffuse interface model of a compressible two-phase flow” In Math. Methods Appl. Sci. 42.5, 2019, pp. 1465–1479 DOI: 10.1002/mma.5436
- [35] E. Feireisl, E. Rocca, E. Petzeltová and G. Schimperna “Analysis of a phase-field model for two-phase compressible fluids” In Math. Models Methods Appl. Sci. 20, 2010, pp. 1129–1160
- [36] Franco Flandoli and Dariusz Gatarek “Martingale and stationary solutions for stochastic Navier-Stokes equations” In Probab. Theory Related Fields 102.3, 1995, pp. 367–391 DOI: 10.1007/BF01192467
- [37] P. J. Flory “Thermodynamics of High Polymer Solutions” In J. Chem. Phys. 10, 1942, pp. 51–61 DOI: 10.1063/1.1723621
- [38] C. G. Gal and M. Grasselli “Longtime behavior for a model of homogeneous incompressible two-phase flows” In Discrete Contin. Dyn. Syst. 28, 2010, pp. 1–39
- [39] C. G. Gal and M. Grasselli “Trajectory attractors for binary fluid mixtures in 3D” In Chinese Ann. Math. Ser. B 31, 2010, pp. 655–678
- [40] M. H. Giga, A. Kirshtein and C. Liu “Variational modeling and complex fluids”, Handbook of mathematical analysis in mechanics of viscous fluids Springer, Cham, 2018
- [41] A. Giorgini “Well-posedness of the two-dimensional Abels-Garcke-Grün model for two-phase flows with unmatched densities” In Calc. Var. Partial Differential Equations 60.3, 2021, pp. Paper No. 100, 40 DOI: 10.1007/s00526-021-01962-2
- [42] A. Giorgini, M. Grasselli and H. Wu “Diffuse interface models for incompressible binary fluids and the mass-conserving Allen-Cahn approximation” In ArXiv e-prints, 2020 arXiv:2005-07236v1 [math.AP]
- [43] A. Giorgini, A. Miranville and R. Temam “Uniqueness and Regularity for the Navier-Stokes-Cahn-Hilliard system” In SIAM J. Math. Anal. 51, 2019, pp. 2535–2574
- [44] A. Giorgini and Roger Temam “Weak and strong solutions to the nonhomogeneous incompressible Navier-Stokes-Cahn-Hilliard system” In J. Math. Pures Appl. (9) 144, 2020, pp. 194–249 DOI: 10.1016/j.matpur.2020.08.009
- [45] A. Giorgini, Roger Temam and Xuan-Truong Vu “The Navier-Stokes-Cahn-Hilliard equations for mildly compressible binary fluid mixtures” In Discrete Contin. Dyn. Syst. Ser. B 26.1, 2021, pp. 337–366 DOI: 10.3934/dcdsb.2020141
- [46] L. Goudenège and L. Manca “Stochastic phase field -Navier-Stokes vesicle-fluid interaction model” In J. Math. Anal. Appl. 496, 2021, pp. 124805
- [47] Ludovic Goudenège “Stochastic Cahn-Hilliard equation with singular nonlinearity and reflection” In Stochastic Process. Appl. 119.10, 2009, pp. 3516–3548 DOI: 10.1016/j.spa.2009.06.008
- [48] Morton E. Gurtin, Debra Polignone and Jorge Viñals “Two-phase binary fluids and immiscible fluids described by an order parameter” In Math. Models Methods Appl. Sci. 6.6, 1996, pp. 815–831 DOI: 10.1142/S0218202596000341
- [49] M. Hairer, K. Lê and T. Rosati “The Allen-Cahn equation with generic initial datum” In ArXiv e-prints, 2022 arXiv:2201.08426 [math.AP]
- [50] M. Hairer, M. Ryser and H. Weber “Triviality of the 2D stochastic Allen-Cahn equation” In Electron. J. Probab. 17, 2012, pp. 1–14
- [51] M. Heida, J. Málek and K. R. Rajagopal “On the development and generalizations of Allen–Cahn and Stefan equations within a thermodynamic framework” In Z. Angew. Math. Phys. 63, 2012, pp. 759–776
- [52] M. Heida, J. Málek and K. R. Rajagopal “On the development and generalizations of Cahn–Hilliard equations within a thermodynamic framework” In Z. Angew. Math. Phys. 63, 2012, pp. 145–169
- [53] P. C. Hohenberg and B. I. Halperin “Theory of dynamic critical phenomena” In Rev. Mod. Phys. 49 American Physical Society, 1977, pp. 435–479 DOI: 10.1103/RevModPhys.49.435
- [54] R. Hošek and V. Mácha “Weak-strong uniqueness for Allen-Cahn/Navier-Stokes system” In Czechoslovak Math. J. 69, 2019, pp. 837–851
- [55] M. L. Huggins “Solutions of Long Chain Compounds” In J. Chem. Phys. 9, 1941, pp. 440 DOI: 10.1063/1.1750930
- [56] Nobuyuki Ikeda and Shinzo Watanabe “Stochastic differential equations and diffusion processes” 24, North-Holland Mathematical Library North-Holland Publishing Co., Amsterdam; Kodansha Ltd., Tokyo, 1989, pp. xvi+555
- [57] J. Jiang, Y.-H. Li and C. Liu “Two-phase incompressible flows with variable density: an energetic variational approach” In Discrete Contin. Dyn. Syst. 37, 2017, pp. 3243–3284
- [58] M. Kotschote “Strong solutions of the Navier-Stokes equations for a compressible fluid of Allen-Cahn type” In Arch. Rational Mech. Anal. 206, 2012, pp. 489–514
- [59] J. A. Langa, J. Real and J. Simon “Existence and Regularity of the Pressure for the Stochastic Navier–Stokes Equations” In Appl. Math. Optim. 48, 2003, pp. 195–210
- [60] Y.-H. Li, S.-J. Ding and M.-X. Huang “Blow-up criterion for an incompressible Navier-Stokes/Allen-Cahn system with different densities” In Discrete Contin. Dyn. Syst. Ser. B 21, 2016, pp. 1507–1523
- [61] Y.-H. Li and M.-X. Huang “Strong solutions for an incompressible Navier-Stokes/Allen-Cahn system with different densities” In Z. Angew. Math. Phys. 69, 2018, pp. Art. 68
- [62] Wei Liu and M. Röckner “Stochastic partial differential equations: an introduction” Springer, Cham, 2015, pp. vi+266 DOI: 10.1007/978-3-319-22354-4
- [63] J. Lowengrub and L. Truskinovsky “Quasi-incompressible Cahn-Hilliard fluids and topological transitions” In Proc. Roy. Soc. Lond. A 454, 1998, pp. 2617–2654
- [64] Carlo Orrieri, Elisabetta Rocca and Luca Scarpa “Optimal control of stochastic phase-field models related to tumor growth” In ESAIM Control Optim. Calc. Var. 26, 2020, pp. 104
- [65] Carlo Orrieri and Luca Scarpa “Singular stochastic Allen-Cahn equations with dynamic boundary conditions” In J. Differential Equations 266.8, 2019, pp. 4624–4667 DOI: 10.1016/j.jde.2018.10.007
- [66] E. Pardoux “Equations aux derivées partielles stochastiques nonlinéaires monotones”, 1975
- [67] M. Röckner, B. Schmuland and X. Zhang “Yamada-Watanabe theorem for stochastic evolution equations in infinite dimensions” In Condens. Matter Phys. 11, 2008, pp. 247–259
- [68] J. Rubinstein and P. Sternberg “Nonlocal reaction-diffusion equations and nucleation” In IMA J. Appl. Math. 48, 1992, pp. 249–264
- [69] M. Ryser, N. Nigam and P. F. Tupper “On the well-posedness of the stochastic Allen–Cahn equation in two dimensions” In J. Comput. Phys. 231.6, 2012, pp. 2537–2550
- [70] L. Scarpa “The stochastic Cahn-Hilliard equation with degenerate mobility and logarithmic potential” In Nonlinearity 34, 2021, pp. 3813
- [71] Luca Scarpa “Analysis and optimal velocity control of a stochastic convective Cahn-Hilliard equation” In J. Nonlinear Sci. 31.2, 2021, pp. Paper No. 45, 57 DOI: 10.1007/s00332-021-09702-8
- [72] Luca Scarpa “On the stochastic Cahn-Hilliard equation with a singular double-well potential” In Nonlinear Anal. 171, 2018, pp. 102–133 DOI: 10.1016/j.na.2018.01.016
- [73] Luca Scarpa “Optimal distributed control of a stochastic Cahn-Hilliard equation” In SIAM J. Control Optim. 57.5, 2019, pp. 3571–3602 DOI: 10.1137/18M1222223
- [74] Luca Scarpa “The stochastic viscous Cahn-Hilliard equation: well-posedness, regularity and vanishing viscosity limit” In Appl. Math. Optim. 84.1, 2021, pp. 487–533 DOI: 10.1007/s00245-020-09652-9
- [75] Jacques Simon “Compact sets in the space ” In Ann. Mat. Pura Appl. (4) 146, 1987, pp. 65–96 DOI: 10.1007/BF01762360
- [76] T. Tachim Medjo “On the convergence of a stochastic 3D globally modified two-phase flow model” In Discrete Contin. Dyn. Syst. 39.1, 2019, pp. 395–430
- [77] T. Tachim Medjo “On the existence and uniqueness of solution to a stochastic 2D Allen-Cahn-Navier-Stokes model” In Stoch. Dyn. 19, 2019, pp. 1950007
- [78] T. Tachim Medjo “On the existence and uniqueness of solution to a stochastic 2D Cahn-Hilliard-Navier-Stokes model” In J. Differential Equations 263, 2017, pp. 1028–1054
- [79] T. Tachim Medjo “On the weak solutions to a stochastic two-phase flow model” In Appl. Anal. 101, 2022, pp. 914–937
- [80] T. Tachim Medjo “On weak martingale solutions to a stochastic Allen-Cahn-Navier-Stokes model with inertial effects” In Discrete Contin. Dyn. Syst. Ser. B, 2021
- [81] Aad W. Vaart and Jon A. Wellner “Weak convergence and empirical processes” With applications to statistics, Springer Series in Statistics Springer-Verlag, New York, 1996, pp. xvi+508 DOI: 10.1007/978-1-4757-2545-2