Asymptotic analysis of linearly elastic flexural shells subjected to an obstacle in absence of friction
Paolo Piersanti
Department of Mathematics and Institute for Scientific Computing and Applied Mathematics, Indiana University Bloomington, 729 East Third Street, Bloomington, Indiana, USA
ppiersan@iu.edu
Abstract.
In this paper we identify a set of two-dimensional variational inequalities governing the displacement of a linearly elastic flexural shell subjected to a confinement condition, expressing that all the points of the admissible deformed configurations remain in a given half-space. The action of friction is neglected.
Keywords. Variational inequalities Flexural shells Penalty method Obstacle problems
MSC 2010. 35J86, 47H07, 74B05.
1. Introduction
Unilateral contact problems arise in many fields such as medicine, engineering, biology and material science. For instance, the description of the motion inside the human heart of the three Aorta valves, which can be regarded as linearly elastic shells, is governed by a mathematical model built up in a way such that each valve remains confined in a certain portion of space without penetrating or being penetrated by the other two valves. In this direction we cite the recent references [40], [41] and [51].
The displacement of a linearly elastic shell is modeled, in general, via the classical equations of three-dimensional linearized elasticity (cf., e.g., [9]). The intrinsic complexity of this model, however, prevents certain situations from being amenably studied like, for instance, when the shell is non-homogeneous and anisotropic (cf., e.g., [7] and [8]), or its thickness varies periodically (cf., e.g., [48] and [49]). It might thus be useful to perform a dimension reduction in order to obtain approximate models which are more amenable to analyze and to implement numerically.
The identification of two-dimensional limit models for time-independent linearly elastic shells was extensively treated by Ciarlet and his associates in the seminal papers [14, 15, 16, 17, 18, 21] for the purpose of justifying Koiter’s model in the case where no obstacles are taken into account. We also mention the papers [44, 45, 47], which are about the numerical computation of the solution of the obstacle-free linear Koiter’s model in the time-dependent case.
For the justification of Koiter’s model in the time-dependent case where the action of temperature is considered, we refer the reader to [33].
In the aforementioned papers no confinement conditions were imposed. The recent papers [38] and [39] provide examples of variational-inequalities-based models arising in biology. The paper [37], instead, treats the modeling and analysis of the melting of a shallow ice sheet by means of a set of doubly nonlinear parabolic variational inequalities.
In the recent papers [19, 20, 22, 23], Ciarlet and his associates fully justified Koiter’s model in the case where the linearly elastic shell under consideration is an elliptic membrane shell subject to the aforementioned confinement condition.
The confinement condition we are here considering considerably departs from the Signorini condition usually considered in the existing literature. Indeed, the Signorini contact condition states that only the “lower face” of the shell is required to remain above the “horizontal” plane (cf., e.g. Chapter 63 of [50]). Such a confinement condition renders the asymptotic analysis considerably more difficult as the constraint now bears on a vector field, the displacement vector field of the reference configuration, instead of on only a single component of this field.
The recovery of a set of two-dimensional variational inequalities as a result of a rigorous asymptotic analysis conducted on a ad hoc three-dimensional model based on the classical equations of three-dimensional linearized elasticity in the case where the linearly elastic shell under consideration is a flexural shell has only been addressed, to our best knowledge, by Léger and Miara in the paper [28] in a very amenable geometrical framework. In the paper [28] the authors do not consider the action of friction. The method proposed in the paper [28], which makes use of a very lengthy argument exploiting the properties of cones, is suitable for treating the case where only one linearly elastic flexural shell is considered. In the context of multi-physics multi-scale problems, like for instance those considered by Quarteroni and his associates in [40], [41] and [51], it is not however clear whether the approach proposed by Léger and Miara is suitable. The purpose of this paper is to remedy this situation by presenting a new method, based on the rigorous asymptotic analysis technique developed by Ciarlet, Lods and Miara [21] and the properties of the penalty method for constrained optimization problem (cf., e.g., [10]), for recovering the same set of two-dimensional variational inequalities for linearly elastic flexural shells that were recovered in the recent papers [19, 20] in a more general geometrical framework. It is also worth mentioning that the method we present in this paper makes use of considerably less lengthy computations than those in [28]. In particular, the approach based on the penalty method we are going to implement will save us the effort of constructing a suitable vector field in the instance of recovering the sought two-dimensional limit model. We refer the reader to the references [26, 32, 43, 46].
More precisely, in this paper we recover, via a rigorous asymptotic analysis as the thickness approaches zero over a ad hoc three-dimensional model (three-dimensional in the sense that it is defined over a three-dimensional subset of ), a set of two-dimensional (two-dimensional in the sense that it is defined over a two-dimensional subset of ) variational inequalities governing the displacement of a linearly elastic flexural shell subject to remain confined in a half-space in absence of friction. The problem under consideration is an obstacle problem.
The paper is divided into five sections (including this one). In section 2 we recall some background and notation. In section 3 we recall the formulation and the properties of a three-dimensional obstacle problem for “general” linearly elastic shell. In section 4 we specialize the formulation presented in the previous section to the case where the linearly elastic shell under consideration is a flexural shell, and then we scale the three-dimensional obstacle problem in a way such that the integration domain becomes independent of the thickness parameter. The penalized version of the three-dimensional obstacle problem is introduced at the end of this section.
Finally, in section 5, a rigorous asymptotic analysis is carried out and the sought set of two-dimensional variational inequalities is recovered.
2. Geometrical preliminaries
For details about the classical notions of differential geometry used in this section and the next one, see, e.g. [11] or [12].
Greek indices, except , take their values in the set , while Latin indices, except when they are used for indexing sequences, take their values in the set , and the summation convention with respect to repeated indices is systematically used in conjunction with these two rules. The notation designates the three-dimensional Euclidean space whose origin is denoted by ; the Euclidean inner product and the vector product of are denoted and ; the Euclidean norm of is denoted . The notation designates the Kronecker symbol.
Given an open subset of , notations such as , , or , designate the usual Lebesgue and Sobolev spaces, and the notation designates the space of all functions that are infinitely differentiable over and have compact supports in . The notation designates the norm in a normed vector space . Spaces of vector-valued functions are denoted with boldface letters.
The positive and negative parts of a function are respectively denoted by:
The boundary of an open subset in is said to be Lipschitz-continuous if the following conditions are satisfied (cf., e.g., Section 1.18 of [13]): Given an integer , there exist constants and , and a finite number of local coordinate systems, with coordinates and , sets , , and corresponding functions
such that
and
We observe that the second last formula takes into account overlapping local charts, while the last set of inequalities express the Lipschitz continuity of the mappings .
An open set is said to be locally on the same side of its boundary if, in addition, there exists a constant such that
Following [13], a domain in is considered as a bounded Lipschitz domain, namely, a bounded and connected open subset of , whose boundary is Lipschitz-continuous, the set being locally on a single side of .
Let be a domain in , let denote a generic point in , and let and . A mapping is an immersion if the two vectors
are linearly independent at each point . Then the image of the set under the mapping is a surface in , equipped with as its curvilinear coordinates. Given any point , the vectors span the tangent plane to the surface at the point , the unit vector
is normal to at , the three vectors form the covariant basis at , and the three vectors defined by the relations
form the contravariant basis at ; note that the vectors also span the tangent plane to at and that .
The first fundamental form of the surface is then defined by means of its covariant components
or by means of its contravariant components
Note that the symmetric matrix field is then the inverse of the matrix field , that and , and that the area element along is given at each point , by , where
Given an immersion , the second fundamental form of the surface is defined by means of its covariant components
or by means of its mixed components
and the Christoffel symbols associated with the immersion are defined by
Given an immersion
and a
vector field , the vector field
can be viewed as a displacement field of the surface , thus defined by means of its covariant components over the vectors of the contravariant bases along the surface. If the norms are small enough, the mapping is also an immersion, so that the set is also a surface in , equipped with the same curvilinear coordinates as those of the surface , called the deformed surface corresponding to the displacement field . One can then define the first fundamental form of the deformed surface by means of its covariant components
The linear part with respect to in the difference is called the linearized change of metric tensor associated with the displacement field , the covariant components of which are thus defined by
The linear part with respect to in the difference is called the linearized change of curvature tensor associated with the displacement field , the covariant components of which are thus defined by
It turns out that, when a generic surface is subjected to a displacement field whose tangential covariant components vanish on a non-zero length portion of boundary of the domain , denoted in the statement of the next result, the following inequality holds (this inequality plays an essential role in our convergence analysis; cf. the proof of Theorem 5.1). Note that the components of the displacement fields, linearized change of metric tensor and linearized change of curvature tensor appearing in the next theorem are no longer assumed to be continuously differentiable functions; they are instead to be understood in a generalized sense, since they now belong to ad hoc Lebesgue or Sobolev spaces.
Throughout the paper the symbol denotes the outer unit normal derivative operator along the boundary .
Theorem 2.1.
Let be a domain in and let an immersion be given. Define the space
Then there exists a constant such that
for all .
∎
The above inequality, which is due to [3] and was later on improved by [4] (see also Theorem 2.6-4 of [11]), constitutes an example of a Korn inequality on a general surface, in the sense that it provides an estimate of an appropriate norm of a displacement field defined on a surface in terms of an appropriate norm of a specific “measure of strain” (here, the linearized change of metric tensor and the linearized change of curvature tensor) corresponding to the displacement field considered.
3. The three-dimensional obstacle problem for a “general” linearly elastic shell
Let be a domain in , let , and let be a non-empty relatively open subset of . For each , we define the sets
we let designate a generic point in the set , and we let . Hence we also have and .
Given an immersion and , consider a shell with middle surface and with constant thickness . This means that the reference configuration of the shell is the set , where the mapping is defined by
One can then show (cf., e.g., Theorem 3.1-1 of [11]) that, if is small enough, such a mapping is an immersion, in the sense that the three vectors
are linearly independent at each point ; these vectors then constitute the covariant basis at the point , while the three vectors defined by the relations
constitute the contravariant basis at the same point. It will be implicitly assumed in the sequel that is small enough so that is an immersion.
One then defines the metric tensor associated with the immersion by means of its covariant components
or by means of its contravariant components
Note that the symmetric matrix field is then the inverse of the matrix field , that and , and that the volume element in is given at each point , , by , where
One also defines the Christoffel symbols associated with the immersion by
Note that .
Given a vector field , the associated vector field
can be viewed as a displacement field of the reference configuration of the shell, thus defined by means of its covariant components over the vectors of the contravariant bases in the reference configuration.
If the norms are small enough, the mapping is also an immersion, so that one can also define the metric tensor of the deformed configuration by means of its covariant components
The linear part with respect to in the difference is then called the linearized strain tensor associated with the displacement field , the covariant components of which are thus defined by
The functions are called the linearized strains in curvilinear coordinates associated with the displacement field .
We assume throughout this paper that, for each , the reference configuration of the shell is a natural state (i.e., stress-free) and that the material constituting the shell is homogeneous, isotropic, and linearly elastic. The behavior of such an elastic material is thus entirely governed by its two Lamé constants and (for details, see, e.g., Section 3.8 of [9]).
We will also assume that the shell is subjected to applied body forces whose density per unit volume is defined by means of its covariant components , and to a homogeneous boundary condition of place along the portion of its lateral face (i.e., the displacement vanishes on ).
For what concerns surface traction forces, the mathematical models characterized by the confinement condition considered in this paper (confinement condition which is also considered in [27] in a more amenable geometrical framework) do not take any surface traction forces into account.
Indeed, there could be no surface traction forces applied to the portion of the three-dimensional shell boundary that engages contact with the obstacle.
The confinement condition considered in this paper is more suitable in the context of multi-scales multi-bodies problems like, for instance, the study of the motion of the human heart valves, conducted by Quarteroni and his associates in [40, 41, 51] and the references therein.
In this paper we consider a specific obstacle problem for such a shell, in the sense that the shell is also subjected to a confinement condition, expressing that any admissible displacement vector field must be such that all the points of the corresponding deformed configuration remain in a half-space of the form
where is a nonzero vector given once and for all. In other words, any admissible displacement field must satisfy
for all , or possibly only for almost all (a.a. in what follows) when the covariant components are required to belong to the Sobolev space as in Theorem 3.1 below.
We will of course assume that the reference configuration satisfies the confinement condition, i.e., that
It is to be emphasized that the above confinement condition considerably departs from the usual Signorini condition favored by most authors, who usually require that only the points of the undeformed and deformed “lower face” of the reference configuration satisfy the confinement condition (see, e.g., [27], [28], [42]). Clearly, the confinement condition considered in the present paper, which is inspired by the formulation proposed by Brézis & Stampacchia [6], is more physically realistic, since a Signorini condition imposed only on the lower face of the reference configuration does not prevent – at least “mathematically” – other points of the deformed reference configuration to “cross” the plane and then to end up on the “other side” of this plane (cf., e.g., Chapter 63 in [50]). It is evident that the vector is thus orthogonal to the plane associated with the half-space where the linearly elastic shell is required to remain confined.
Such a confinement condition renders the asymptotic analysis considerably more difficult, however, as the constraint now bears on a vector field, the displacement vector field of the reference configuration, instead of on only a single component of this field.
The mathematical modeling of such an obstacle problem for a linearly elastic shell is then clear; since, apart from the confinement condition, the rest, i.e., the function space and the expression of the quadratic energy , is classical (see, e.g. [11]). More specifically, let
denote the contravariant components of the elasticity tensor of the elastic material constituting the shell. Then the unknown of the problem, which is the vector field where the functions are the three covariant components of the unknown “three-dimensional” displacement vector field of the reference configuration of the shell, should minimize the energy defined by
for each
over the set of admissible displacements defined by:
The solution to this minimization problem exists and is unique, and it can be also characterized as the unique solution of the following problem:
Problem .
Find that satisfies the following variational inequalities:
for all .
The following result can be thus straightforwardly proved.
Theorem 3.1.
The quadratic minimization problem: Find a vector field such that
has one and only one solution. Besides, is also the unique solution of Problem .
Proof.
Define the space
Then, thanks to the uniform positive-definiteness of the elasticity tensor [9], and to the boundary condition of place satisfied on (recall that , and that is a non-empty relatively open subset of ), it can be shown (see Theorems 3.8-3 and 3.9-1 of [12]) that the continuous and symmetric bilinear form
is -elliptic; besides, the linear form
is clearly continuous. Finally, the set is nonempty (by assumption), closed in (any convergent sequence in contains a subsequence that pointwise converges almost everywhere to its limit), and convex (as is immediately verified).
The existence and uniqueness of the solution to the minimization problem and its characterization by means of variational inequalities is then classical (see, e.g., [13], [24] or [25]).
∎
Since , it evidently follows that for all . But in fact, a stronger property holds (cf. Lemma 2.1 of [23]).
Lemma 3.1.
Let be a domain in , let be an immersion, let be a non-zero vector, and let . Then the inclusion
implies that
∎
We now consider the “penalized” version of Problem . One such penalization transforms the set of variational inequalities in Problem into a set of nonlinear variational equations, where the nonlinearity is defined in terms of the measure of the “violation” of the constraint.
Let denote a penalty parameter. The “penalized” variational formulation corresponding to Problem takes the following form:
Problem .
Find that satisfies the following variational equations:
for all .
Note that the penalty term corresponds to the operator defined by:
(1)
Note that the denominator appearing in the definition of is always positive, since the three vector fields are linearly independent at each .
In order to show the existence and uniqueness of solution for Problem , we have to show that the operator is monotone. The following lemma, whose proof can be found, for instance, in [37] serves for this purpose.
Lemma 3.2.
Let , with an integer, be an open set. The operator defined by
is monotone, bounded and Lipschitz continuous with Lipschitz constant equal to .
∎
The existence and uniqueness of the solution for Problem is classical too (cf., e.g., Theorem 3.15 in [26], or [29]).
Theorem 3.2.
Problem has a unique solution.
Proof.
Observe that the linear form
is continuous (cf. Theorem 3.1). The mapping defined by
is continuous and, thanks to an ad hoc Korn’s inequality in curvilinear coordinates (cf., e.g., Theorem 1.7-4 of [11]) and Lemma 3.2, strictly monotone and coercive. Therefore, the conclusion follows by the Minty-Browder theorem (cf., e.g., Theorem 9.14-1 of [13]).
∎
By means of a different classical approach based on energy estimates [29, 43] (see also part (iv) of Theorem 9.14-1 of [13]), one can prove Theorem 3.2 by establishing that for each
(2)
and that, thanks to the monotonicity of the penalty term established in Lemma 3.2, the weak limit satisfies Problem . As a result of the convergence (2) and the continuity of the linear operator , we have that
Actually, a stronger conclusion holds: Given , the following strong convergence holds
To see this, observe that the uniform positive-definiteness of the elasticity tensor (cf., e.g., [9]), Korn’s inequality (viz. Theorem 1.7-4 of [11]), the monotonicity of the penalty term (Lemma 3.2), the continuity of the operators , and the weak convergence (2) give
as .
In particular, we have that for each and for each , we can find a number such that, for each , it results
(3)
for all , where and respectively denote the solutions of Problem and Problem .
We observe that if enjoys higher regularity (viz., e.g., [43] and [30]) then the threshold number can be made independent of .
Theorem 3.3.
Let be given.
Let be the solution of Problem and let be the solution of Problem .
Assume that there exists a constant independent of and such that , where the nonlinear operator has been defined in (1).
Then, there exists a constant independent of and such that:
Proof.
For each , define the operator by:
Define , and observe that . Indeed, a direct computation gives
Let us estimate
where the latter inequality holds thanks to the assumed estimate for the penalty term. Since , an application of the uniform positive definiteness of the fourth order three-dimensional elasticity tensor (Theorem 1.8-1 of [11]), Korn’s inequality (Theorem 1.7-4 in [11]) gives
where we recall that and are positive constants independent of (viz., e.g., Theorem 3.1-1 of [11]) and that the bounding constant for the fourth order three-dimensional elasticity tensor is independent of (viz., e.g., Theorem 3.3-2 of [11]).
In conclusion, we have that
so that
where the constant is defined by
This completes the proof.
∎
The property that there exists a constant independent of and such that , where has been defined in (1) could be established upon proving, in the same spirit of [43] (see also [30]), that the solution of Problem is of class and that the solution is also of class . However, the latter augmentations of regularity are not easy at all to prove, as the boundary conditions for the corresponding problems are only enforced on a portion of the boundary, thus preventing us from applying the argument of Agmon, Douglis & Nirenberg [1, 2].
4. The scaled three-dimensional problem for a family of flexural shells
In section 3, we considered an obstacle problem for “general” linearly elastic shells. From now on, we will restrict ourselves to a specific class of shells, according to the following definition (proposed in [21]; see also [11]).
Consider a linearly elastic shell, subjected to the various assumptions set forth in section 3. Such a shell is said to be a linearly elastic flexural shell if the following two additional assumptions are satisfied: first, , i.e., the homogeneous boundary condition of place is imposed over a nonzero area portion of the entire lateral face of the shell, and second, the space
contains nonzero functions, i.e., .
In this paper, we consider the obstacle problem as defined in section 3for a family of linearly elastic flexural shells, all sharing the same middle surface and whose thickness is considered as a “small” parameter approaching zero. Our objective then consists in performing an asymptotic analysis as , so as to seek whether we can identify a limit two-dimensional problem. To this end, we shall resort to a (by now standard) methodology first proposed by Ciarlet, Lods and Miara (cf. Theorem 5.1 of [21] and Theorem 6.2-1 of [11]): To begin with, we “scale” each problem , over a fixed domain , using appropriate scalings on the unknowns and assumptions on the data. Note that these scalings and assumptions definitely depend on the type of shells that are considered; for instance, those used for the linearly elastic elliptic membrane shells considered elsewhere (cf. [22] and also [23]) are different.
More specifically, let
let denote a generic point in the set , and let . With each point , we associate the point defined by
so that and . To the unknown and to the vector fields appearing in the formulation of Problem corresponding to a linearly elastic flexural shell, we then associate the scaled unknown and the scaled vector fields by letting
at each . Finally, we assume that there exist functions independent on such that the following assumptions on the data hold:
Note that the independence on of the Lamé constants assumed in section 3 in the formulation of Problem implicitly constituted another assumption on the data.
In view of the proposed scaling, we define the “scaled” version of the geometrical entities introduced in section 2:
where
Define the space
and define, for each , the set
We are thus in a position to introduce the “scaled” version of Problem , that will be denoted in what follows by .
Problem .
Find that satisfies the following variational inequalities:
for all .
Theorem 4.1.
The scaled unknown is the unique solution of the variational Problem .
Proof.
The variational Problem simply constitutes a re-writing of the variational Problem , this time in terms of the scaled unknown , of the vector fields , and of the functions , which are now all defined over the domain . Then the assertion follows from this observation.
∎
The functions appearing in Problem are called the scaled linearized strains in curvilinear coordinates associated with the scaled displacement vector field .
For later purposes (like in Lemma 4.1 below), we also let
Likewise, one can introduce the “scaled” version of Problem , that will be denoted in what follows by .
Problem .
Find that satisfies the following variational equations:
for all .
The following existence and uniqueness result can be thus easily proved in the same fashion as Theorem 3.2.
Theorem 4.2.
The scaled unknown is the unique solution of the variational Problem .
Proof.
The variational Problem simply constitutes a re-writing of the variational Problem , this time in terms of the scaled unknown , of the vector fields , and of the functions , which are now all defined over the domain . Then the assertion follows from this observation.
∎
By means of an analogous reasoning, a condition similar to (3) can be derived, i.e., for each we can find a number such that, for each , it results
(4)
for each , where and respectively denote the solutions of Problem and Problem .
In the same spirit as Theorem 3.3 it can be shown that if the scaled solution enjoys higher regularity then the threshold value can be made independent of .
Without loss of generality (cf., e.g., [43]), given any , we restrict ourselves to considering penalty parameters with the following property:
(5)
It is straightforward to observe that as .
We observe that the variational Problem could have been equivalently written as a minimization problem, thus mimicking that found in Theorem 3.1. It turns out, however, that its formulation in Theorem 4.2 as a set of penalized variational equations is more convenient for the asymptotic analysis undertaken in section 5.
It is immediately verified (cf., e.g., [11]) that other assumptions on the data are possible that would give rise to the same problem over the fixed domain . For instance, should the Lamé constants (now denoted) and appearing in Problem be of the form and , where and are constants independent of and is an arbitrary real number, the same Problem arises if we assume that the components of the applied body force density are now of the form
where the functions are independent of .
The next lemma assembles various asymptotic properties as of functions and vector fields appearing in the formulation of Problem ; these properties will be repeatedly used in the proof of the convergence theorem (Theorem 5.1).
In the statement of the next preparatory lemma (cf., e.g., Theorems 3.3-1 and 3.3-2 of [11]), the notation “”, or “”, stands for a remainder that is of order , or , with respect to the sup-norm over the set , and any function, or vector-valued function, of the variable , such as , etc. (all these are defined in section 2) is identified with the function, or vector-valued function, of that takes the same value at and is independent of ; for brevity, this extension from to is designated with the same notation.
Recall that is implicitly assumed to be small enough so that is an immersion.
Lemma 4.1.
Let be an immersion.
Let be defined as in Theorem 3.1-1 of [11]. The functions have the following properties:
for all , where
and there exists a constant such that
for all , all , and all symmetric matrices .
The functions and have the following properties:
for all and all . In particular then, there exist constants and such that
The vector fields and have the following properties:
∎
We recall (cf., e.g., [11]), that the various relations and estimates in Lemma 4.1 hold in fact for any family of linearly elastic shells, i.e., irrespective of whether these shells are flexural ones or not.
When one considers a family of linearly elastic flexural shells whose thickness approaches zero, a specific Korn’s inequality in curvilinear coordinates (cf., e.g., Theorem 4.1 of [21] or Theorem 5.3-1 of [11]) holds over the fixed domain , according to the following theorem. That the constant that appears in this inequality is independent of plays a key role in the asymptotic analysis of such a family (see part (i) of the proof of Theorem 5.1).
Theorem 4.3.
Let be an immersion.
Let there be given a family of linearly elastic flexural shells with the same middle surface and thickness . Define the space
Then there exist constants and such that
for all and all .
∎
5. Rigorous asymptotic analysis
The ultimate goal of this paper is to show, in the same spirit as [21] (see also Theorem 6.2-1 of [11]), that the solutions of the (scaled) three-dimensional problems converge - as approaches zero - to the solution of a two-dimensional problem, denoted by in what follows.
The proof proposed in [21] (see also Theorem 6.2-1 of [11]) resorts, however, to the usage of a specific vector field that was first introduced by Miara and Sanchez-Palencia [31]. This construction argument is, in general, not applicable to the context of variational inequalities, for which the test functions are chosen in a nonempty, closed and convex subset of a certain space.
In order to overcome this difficulty, we first prove that, under the assumption (5), the solutions of Problem converge - as approaches zero - to the solution of the variational problem .
Define the set
where the space has been defined in section 4. We observe that the set is nonempty, closed and convex in the space . The vector fields and the functions , and , have been defined in section 2.
We are thus in a position to define the two-dimensional problem as follows:
Problem .
Find that satisfies the following variational inequalities:
for all , where
In the same spirit as Theorem 3.1, it can be show that Problem admits one and only one solution.
We are now ready to show that, under the assumption (5), the solutions of Problem converge - as approaches zero - to the solution of Problem .
Differently from the linearly elastic elliptic membrane case, we will see that problems arise when one has to define an appropriate test vector field for recovering the limit model, which is the main objective of Theorem 5.1 (cf. part (iv)). Since this test vector field will have to depend on the partial derivatives of the displacement, it is hard to find an expression for it that takes into account the geometrical constraint as well. For this reason, the penalty method seems to be the most convenient technique to attack this problem.
Theorem 5.1.
Let be a domain in , let be the middle surface of a flexural shell, let be a non-zero length portion of the boundary (cf. section 4) and let be a non-zero vector given once and for all. Let us consider the non-trivial space (cf. section 4)
and let us define the set
Let there be given a family of linearly elastic flexural shells with the same middle surface and thickness , and let denote for each the unique solution of Problem , where the penalty parameter is assumed to be as in (5).
Then there exists independent of the variable and satisfying
Define the average
Then
where is the unique solution to the two-dimensional variational Problem .
Proof.
Strong and weak convergences as are respectively denoted by and . For brevity, we let
The outline of the proof, which is broken into six parts numbered (i)–(vi), is to a large extent inspired by the proof of Theorem 6.2-1 of [11] (itself adapted from Theorem 5.1 in Ciarlet, Lods and Miara [21]), where no confinement condition was imposed. This is why some parts of the proof are reminiscent of those in [11]; otherwise, considering the confinement condition requires extra care.
(i) There exists a subsequence, still denoted , and there exists and there exist satisfying
and such that
Letting in the variational equations of Problem . Combining the uniform positive-definiteness of the tensor , the Korn inequality of Theorem 4.3, the asymptotic behavior of the function (Lemma 4.1), and the fact that (see Lemmas 3.1 and 3.2)
we obtain for sufficiently small:
This chain of inequalities first shows that the norms are bounded independently of , secondly, that the terms are bounded independently of and, finally, that the terms
(6)
are bounded independently of as well (recall that, by assumption (5), we have that ).
Hence there exist a subsequence, a vector field , and functions such that, if we let , the following convergence process occurs:
(7)
The fact that is a consequence of the Rellich-Kondrašov Theorem (viz., e.g., Theorem 6.6-3 of [13]). Note that, by the definition of limit, the latter means that for any given there exists a number such that for all it results:
(8)
Note that the third convergence in (7) means that for any given there exists a number such that for all it results:
(9)
Moreover, recall that the term (6) is bounded independently of , in the sense that there exists a constant such that:
(10)
Therefore, it suffices to take .
Note that the fourth convergence in (7) means that for any given there exists a number such that for all it results:
(11)
Moreover, recall that the term (6) is bounded independently of , in the sense that there exists a constant such that:
(12)
so that is a suitable threshold number for verifying the definition of limit.
Combining the convergence (8) with the fourth convergence in (7), the continuity of the negative part operator (Lemma 3.2), and the fact that, for a generic function,
gives:
That on follows from the continuity of the trace operator . Indeed, for all , we have that for all such that on ,
so that a density result proved by Bernard [5] (see also Theorem 6.7-3 of [13]) gives on .
(ii) The weak limits found in (i) are independent of the variable , in the sense that they satisfy
Besides, the average satisfies , namely,
Apart from the latter property, the proof is identical to that of part (ii) of the proof of Theorem 6.2-1 in [11]. Let us thus prove that
By part (i), we have that
Since is independent of , we have that an application of Theorem 4.2-1 (a) of [11] and part (i) gives
so that .
(iii) The weak limits and found in (i) satisfy
The equality in follows from Theorem 5.2-2 of [11].
Let be arbitrarily chosen. It is known (cf., e.g., part (iii) of Theorem 6.2-1 of [11]) that
These relations, combined with the boundedness of the terms independently of (part (i)), the asymptotic behavior of the functions , the third convergence in (7), and as (Lemma 4.1), give
Consequently,
Since this inequality holds for any vector field , it follows by Theorem 3.4-1 of [11] that
In particular, the latter gives:
(iv) The weak limit is the unique solution of Problem .
Given any , define the vector field by:
(13)
We have that and for all .
Following the same steps as in part (iv) of [11], observe that:
(14)
Fix and define the number
so that, combining the uniform positive-definiteness of the three-dimensional elasticity tensor and the asymptotic behavior of the function (Lemma 4.1), we obtain:
An application of the latter to the variational equations of Problem under the specialization , with gives:
(15)
By virtue of the definition of the vector field introduced in (13) and the asymptotic behaviour of the contravariant basis vectors established in Lemma 4.1, we obtain that the term
(16)
has less or equal than zero when since the first addend in the last equality is less or equal than zero for all being , the second and the third addends tend to zero as thanks to the third convergence of (7), and the fourth addend is less or equal than zero thansk to the monotonicity of established in Lemma 3.2.
Note that the second factor in the second integral of the last equality corresponds to a remnant of the definition of (cf. Lemma 4.1).
The asymptotic behavior of the functions and exhibited in (14), the asymptotic behavior of the three-dimensional elasticity tensor and (Lemma 4.1), the weak convergences in established in part (i), and the relations satisfied by (part (iii)) together give:
(17)
We have yet to take into account the relations in established in part (iii). Since is independent of (part (ii)), these relations show that the functions are of the form
The asymptotic behaviors observed in (15)–(17), Lemma 3.1, and the uniform positive-definiteness of the fourth order two-dimensional elasticity tensor in turn imply:
In conclusion, the latter yields
which establishes that is the unique solution for Problem since is arbitrarily chosen.
(v) The weak convergence in established in part (i) is in fact strong, i.e.,
and holds for the whole sequence .
The proof is identical to that of part (vi) of the proof of Theorem 6.2-1 of [11] and for this reason is omitted. The assertion means that for all there exists a number such that for all it results
(vi) The weak convergences in established in part (i) are in fact strong, i.e.,
Besides, the limits are unique; hence these convergences hold for the whole family .
Let be as in part (iv). Since , we define a vector field as follows:
Specializing in the variational equations of Problem and repeating the same computations as in part (iv) gives:
In conclusion, by the uniform positive-definiteness of the two-dimensional fourth-order elasticity tensor (Lemma 4.1) and the asymptotic behavior of the sequence of vector fields in (18), we obtain and that
These relations in turn imply that the strong convergence
holds. The functions are uniquely determined, since they are given by
and the vector field is uniquely determined as the unique solution of Problem . That the functions are uniquely determined then follows from the relations established in part (iii). Therefore, the whole sequence strongly converges to the functions in and the proof is complete.
∎
Observe that the conclusion of Theorem 5.1 and Theorem 4.2-1(b) of [11] give that for each and for each there exists a number such that for all it results
(19)
where is the unique solution of Problem , and is the unique solution of Problem , and is as in (5).
Therefore, the convergence as of the solutions of Problem to the solution of Problem can thus be established as a direct corollary of Theorem 3.3 and Theorem 5.1.
Corollary 5.2.
Let be a domain in , let be the middle surface of a flexural shell, let be a non-zero length portion of the boundary (cf. section 4) and let be a non-zero vector given once and for all. Let us consider the non-trivial space (cf. section 4)
and let us define the set
Let there be given a family of linearly elastic flexural shells with the same middle surface and thickness , and let denote, for each , the unique solution of Problem .
Then, we have that
where is the unique solution to the two-dimensional variational Problem .
Proof.
To prove the claim, we have to show that for each and for all there exists a number such that for all it results
(20)
where is the unique solution of Problem and is the unique solution of Problem .
In order to prove (20), fix and take . For each we have that an application of (4), the triangle inequality, Theorem 4.2-1 (b) of [11], and (19) gives
(21)
whenever (viz. Theorem 3.3 for a reasonable sufficient condition ensuring that is independent of ) and is as in (5).
Observe that if we take as in (5) and it resulted , then it would suffice to reduce until the requirement for the penalty parameter is met. In doing so, the fact that is independent of plays a critical role in establishing the sought convergence.
The estimate (21) means that the average across the thickness of the solutions of Problem converge to the solution of Problem as , as it had to be proved.
∎
Figure 1. Geometrical interpretation of Corollary 5.2 and criticality of the condition (5).
Let be fixed and let , where is the threshold parameter found in Corollary 5.2.
Let denote the solution to Problem .
Let be a “penalty” parameter such that , and for which the distance between the averaged solution of Problem (i.e., the vector field ) and the solution of Problem with respect to the norm is less than .
If , it might happen, as a consequence of (4), that the distance between the averaged solution of Problem (i.e., the vector field ) and the averaged solution of Problem (i.e., the vector field ) is, with respect to the norm, strictly greater than the fixed number .
If, however, we further reduce the parameter until , then we can find an element , solution of Problem , whose distance (once again with respect to the norm) from both and is strictly less than . In the latter step, a critical role is played by the fact that is assumed to be independent of (viz. Theorem 3.3 for a reasonable sufficient condition). This figure originally appeared in [36].
It remains to “de-scale” the results of Theorem 5.1 and Corollary 5.2, which apply to the solutions of the scaled problem . This means that we need to convert these results into ones about the unknown , which represents the physical three-dimensional vector field of the actual reference configuration of the shell. As shown in the next theorem, this conversion is most conveniently achieved through the introduction of the averages across the thickness of the shell.
Theorem 5.3.
Let the assumptions on the data be as in section 4 and let the assumptions on the immersion be as in Theorem 4.3. Let denote for each the unique solution of the variational Problem and let denote the unique solution to the variational inequalities in Problem . Then
Proof.
The proof is analogous to that of Theorem 6.4-1 in [11] and for this reason is omitted.
∎
In view of the scalings, viz., at each , and of the assumption on the data, viz., at each , made in section 4, it is natural to also “de-scale” the unknown appearing in the limit two-dimensional problem found in Theorem 4.3, by letting
and by using in its formulation the contravariant components
instead of their scaled counterparts . In this fashion, it is immediately found that is the unique solution to the variational inequalities
for all . These inequalities now display the factor , which always appears in the left-hand sides of equations modelling flexural shells.
Conclusion and final remarks
In this paper we identified a set of two-dimensional variational inequalities that model the displacement of a linearly elastic flexural shell subjected to a confinement condition, expressing that all the points of the admissible deformed configurations remain in a given half-space.
The starting point of the rigorous asymptotic analysis we carried out is a set of variational inequalities based on the classical equations of three-dimensional linearized elasticity, and posed over a non-empty, closed and convex subset of a suitable Sobolev space. These variational inequalities govern the displacement of a three-dimensional flexural shell subject to a confinement condition like the one recalled beforehand.
By means of the penalized version of the aforementioned problem (i.e., the set of variational inequalities based on the classical equations of three-dimensional linearized elasticity), we managed to show that, as the thickness parameter approaches zero, the average across the thickness of the solution of the original three-dimensional model converges to the solution of a ad hoc two dimensional model. In this regard, it is worth recalling that the rigorous asymptotic analysis (Theorem 5.1) hinged on an ad hoc scaling of the penalty coefficient with respect to the shell thickness as well as the critical assumption (5).
The two-dimensional model we recovered in this paper coincides with the two-dimensional model recovered as a result of a rigorous asymptotic analysis carried out starting from Koiter’s model in the case where the linearly elastic shell under consideration is subjected to an obstacle (viz. [19] and [20]).
It is worth mentioning that, unlike the case where the linearly elastic shell under consideration was a linearly elastic elliptic membrane shell (viz. [22], [23] and [36]), the rigorous asymptotic analysis carried out in this paper does not resort to any additional assumption apart from those already made in [21] (see also Chapter 6 of [11]).
Indeed, when the linearly elastic shell under consideration was a linearly elastic elliptic membrane shell, the authors had to resort to a ad hoc “density property” in order to carry out the asymptotic analysis leading to the identification of a suitable two-dimensional set of variational inequalities.
The “density property” we mentioned is ensured only under certain geometrical assumptions, which restrict the applicability of the recalled result. This very “density property” was recently exploited in [35] (see also [34]) to show that the solution of a certain obstacle problem in linearized elasticity enjoys, at least locally, higher regularity properties.
Acknowledgements
The author is greatly indebted to Professor Philippe G. Ciarlet for his encouragement and guidance.
The author would like to express his sincere gratitude to the Anonymous Referees for their suggested improvements.
The author was partly supported by the Research Fund of Indiana University.
The author declares that no experimental data was used in the preparation of this manuscript.
The author declares that there is no conflict of interests.
References
Agmon et al. [1959]
S. Agmon, A. Douglis, and L. Nirenberg.
Estimates near the boundary for solutions of elliptic partial
differential equations satisfying general boundary conditions. I.
Comm. Pure Appl. Math., 12:623–727, 1959.
Agmon et al. [1964]
S. Agmon, A. Douglis, and L. Nirenberg.
Estimates near the boundary for solutions of elliptic partial
differential equations satisfying general boundary conditions. II.
Comm. Pure Appl. Math., 17:35–92, 1964.
Bernadou and Ciarlet [1976]
M. Bernadou and P. G. Ciarlet.
Sur l’ellipticité du modèle linéaire de coques de W. T.
Koiter.
Computing methods in applied sciences and engineering (Second
Internat. Sympos., Versailles, 1975), Part 1. Lecture Notes in
Econom. and Math. Systems, 134:89–136, 1976.
Bernadou et al. [1994]
M. Bernadou, P. G. Ciarlet, and B. Miara.
Existence theorems for two-dimensional linear shell theories.
J. Elasticity, 34:111–138, 1994.
Bernard [2011]
J. M. E. Bernard.
Density results on Sobolev spaces whose elements vanish on a part
of the boundary.
Chinese Ann. Math., Ser B, 32:823–846, 2011.
Brezis and Stampacchia [1968]
H. Brezis and G. Stampacchia.
Sur la régularité de la solution d’inéquations
elliptiques.
Bull. Soc. Math. France, 96:153–180, 1968.
Caillerie and
Sánchez-Palencia [1995a]
D. Caillerie and E. Sánchez-Palencia.
A new kind of singular stiff problems and application to thin elastic
shells.
Math. Models Methods Appl. Sci., 5(1):47–66, 1995a.
Caillerie and
Sánchez-Palencia [1995b]
D. Caillerie and E. Sánchez-Palencia.
Elastic thin shells: asymptotic theory in the anisotropic and
heterogeneous cases.
Math. Models Methods Appl. Sci., 5(4):473–496, 1995b.
Ciarlet [1988]
P. G. Ciarlet.
Mathematical Elasticity. Vol. I: Three-Dimensional Elasticity.
North-Holland, Amsterdam, 1988.
Ciarlet [1989]
P. G. Ciarlet.
Introduction to Numerical Linear Algebra and Optimisation.
Cambridge University Press, 1989.
Ciarlet [2000]
P. G. Ciarlet.
Mathematical Elasticity. Vol. III: Theory of Shells.
North-Holland, Amsterdam, 2000.
Ciarlet [2005]
P. G. Ciarlet.
An Introduction to Differential Geometry with
Applications to Elasticity.
Springer, Dordrecht, 2005.
Ciarlet [2013]
P. G. Ciarlet.
Linear and Nonlinear Functional Analysis with Applications.
Society for Industrial and Applied Mathematics, Philadelphia, 2013.
Ciarlet and Destuynder [1979]
P. G. Ciarlet and P. Destuynder.
A justification of the two-dimensional linear plate model.
J. Mécanique, 18:315–344, 1979.
Ciarlet and Lods [1996a]
P. G. Ciarlet and V. Lods.
On the ellipticity of linear membrane shell equations.
J. Math. Pures Appl., 75:107–124,
1996a.
Ciarlet and Lods [1996b]
P. G. Ciarlet and V. Lods.
Asymptotic analysis of linearly elastic shells. I. Justification
of membrane shell equations.
Arch. Rational Mech. Anal., 136(2):119–161, 1996b.
Ciarlet and Lods [1996c]
P. G. Ciarlet and V. Lods.
Asymptotic analysis of linearly elastic shells. III.
Justification of koiter’s shell equations.
Arch. Rational Mech. Anal., 136(2):191–200, 1996c.
Ciarlet and Lods [1996d]
P. G. Ciarlet and V. Lods.
Asymptotic analysis of linearly elastic shells: “generalized
membrane shells”.
J. Elasticity, 43(2):147–188,
1996d.
Ciarlet and Piersanti [2019a]
P. G. Ciarlet and P. Piersanti.
Obstacle problems for Koiter’s shells.
Math. Mech. Solids, 24:3061–3079,
2019a.
Ciarlet and Piersanti [2019b]
P. G. Ciarlet and P. Piersanti.
A confinement problem for a linearly elastic Koiter’s shell.
C.R. Acad. Sci. Paris, Sér. I, 357:221–230,
2019b.
Ciarlet et al. [1996]
P. G. Ciarlet, V. Lods, and B. Miara.
Asymptotic analysis of linearly elastic shells. II. Justification
of flexural shell equations.
Arch. Rational Mech. Anal., 136(2):163–190, 1996.
Ciarlet et al. [2018]
P. G. Ciarlet, C. Mardare, and P. Piersanti.
Un problème de confinement pour une coque membranaire
linéairement élastique de type elliptique.
C.R. Acad. Sci. Paris, Sér. I, 356(10):1040–1051, 2018.
Ciarlet et al. [2019]
P. G. Ciarlet, C. Mardare, and P. Piersanti.
An obstacle problem for elliptic membrane shells.
Math. Mech. Solids, 24(5):1503–1529, 2019.
Duvaut and Lions [1976]
G. Duvaut and J. L. Lions.
Inequalities in Mechanics and Physics.
Springer, Berlin, 1976.
Glowinski [1984]
R. Glowinski.
Numerical Methods for Nonlinear Variational Problems.
Springer-Verlag, New York, 1984.
Kikuchi and Oden [1988]
N. Kikuchi and J. T. Oden.
Contact problems in elasticity: a study of variational
inequalities and finite element methods, volume 8.
Society for Industrial and Applied Mathematics (SIAM), Philadelphia,
PA, 1988.
Léger and Miara [2008]
A. Léger and B. Miara.
Mathematical justification of the obstacle problem in the case of a
shallow shell.
J. Elasticity, 90:241–257, 2008.
Léger and Miara [2018]
A. Léger and B. Miara.
A linearly elastic shell over an obstacle: The flexural case.
J. Elasticity, 131:19–38, 2018.
Lions [1969]
J. L. Lions.
Quelques méthodes de résolution des problèmes aux
limites non linéaires.
Dunod; Gauthier-Villars, Paris, 1969.
Meixner and Piersanti [Submitted]
A. Meixner and P. Piersanti.
Numerical approximation of the solution of an obstacle problem
modelling the displacement of elliptic membrane shells via the penalty
method.
Submitted.
URL https://arxiv.org/abs/2205.11293.
Miara and Sanchez-Palencia [1996]
B. Miara and E. Sanchez-Palencia.
Asymptotic analysis of linearly elastic shells.
Asymptotic Anal., 12(1):41–54, 1996.
Piersanti [2020]
P. Piersanti.
An existence and uniqueness theorem for the dynamics of flexural
shells.
Math. Mech. Solids, 25(2):317–336, 2020.
Piersanti [2021]
P. Piersanti.
On the justification of the frictionless time-dependent Koiter’s
model for thermoelastic shells.
J. Differential Equations, 296:50–106, 2021.
Piersanti [2022a]
P. Piersanti.
On the improved interior regularity of the solution of a fourth order
elliptic problem modelling the displacement of a linearly elastic shallow
shell lying subject to an obstacle.
Asymptot. Anal., 127(1-2):35–55,
2022a.
Piersanti [2022b]
P. Piersanti.
On the improved interior regularity of the solution of a second order
elliptic boundary value problem modelling the displacement of a linearly
elastic elliptic membrane shell subject to an obstacle.
Discrete Contin. Dyn. Syst., 42(2):1011–1037, 2022b.
Piersanti [2022c]
P. Piersanti.
Asymptotic analysis of linearly elastic elliptic membrane shells
subjected to an obstacle.
J. Differential Equations, 320:114–142,
2022c.
Piersanti and Temam [2023]
P. Piersanti and R. Temam.
On the dynamics of grounded shallow ice sheets: Modelling and
analysis.
Adv. Nonlinear Anal., 12(1):40 pp., 2023.
Piersanti et al. [2022a]
P. Piersanti, K. White, B. Dragnea, and R. Temam.
Modelling virus contact mechanics under atomic force imaging
conditions.
Appl. Anal., 101(11):3947–3957,
2022a.
Piersanti et al. [2022b]
P. Piersanti, K. White, B. Dragnea, and R. Temam.
A three-dimensional discrete model for approximating the deformation
of a viral capsid subjected to lying over a flat surface in the static and
time-dependent case.
Anal. Appl. (Singap.), 20(6):1159–1191, 2022b.
Piersanti et al. [2021]
R. Piersanti, P. C. Africa, M. Fedele, C. Vergara, L. Dedè, A. F. Corno,
and A. Quarteroni.
Modeling cardiac muscle fibers in ventricular and atrial
electrophysiology simulations.
Comput. Methods Appl. Mech. Engrg., 373:113468, 33,
2021.
Regazzoni et al. [2021]
F. Regazzoni, L. Dedè, and A. Quarteroni.
Active force generation in cardiac muscle cells: mathematical
modeling and numerical simulation of the actin-myosin interaction.
Vietnam J. Math., 49(1):87–118, 2021.
Rodríguez-Arós [2018]
A. Rodríguez-Arós.
Mathematical justification of the obstacle problem for elastic
elliptic membrane shells.
Applicable Anal., 97:1261–1280, 2018.
Scholz [1984]
R. Scholz.
Numerical solution of the obstacle problem by the penalty method.
Computing, 32(4):297–306, 1984.
Shen and Li [2018]
X. Shen and H. Li.
The time-dependent Koiter model and its numerical computation.
Appl. Math. Model., 55:131–144, 2018.
Shen et al. [2019]
X. Shen, J. Jia, S. Zhu, H. Li, L. Bai, T. Wang, and X. Cao.
The time-dependent generalized membrane shell model and its numerical
computation.
Comput. Methods Appl. Mech. Engrg., 344:54–70,
2019.
Shen et al. [2020a]
X. Shen, L. Piersanti, and P. Piersanti.
Numerical simulations for the dynamics of flexural shells.
Math. Mech. Solids, 25(4):887–912,
2020a.
Shen et al. [2020b]
X. Shen, Q. Yang, L. Li, Z. Gao, and T. Wang.
Numerical approximation of the dynamic Koiter’s model for the
hyperbolic parabolic shell.
Appl. Numer. Math., 150:194–205,
2020b.
Telega and Lewiński [1998a]
J. J. Telega and T. Lewiński.
Homogenization of linear elastic shells: -convergence and
duality. Part I. Formulation of the problem and the effective model.
Bull. Polish Acad. Sci., Techincal Sci., 46:1–9,
1998a.
Telega and Lewiński [1998b]
J. J. Telega and T. Lewiński.
Homogenization of linear elastic shells: -convergence and
duality. Part II. Dual homogenization.
Bull. Polish Acad. Sci., Techincal Sci., 46:11–21,
1998b.
Zeidler [1988]
E. Zeidler.
Nonlinear functional analysis and its applications. IV.
Springer-Verlag, New York, 1988.
Applications to mathematical physics, Translated from the German and
with a preface by Juergen Quandt.
Zingaro et al. [2021]
A. Zingaro, L. Dedè, F. Menghini, and A. Quarteroni.
Hemodynamics of the heart’s left atrium based on a Variational
Multiscale-LES numerical method.
Eur. J. Mech. B Fluids, 89:380–400, 2021.