Boundary-layer analysis of repelling particles pushed to an impenetrable barrier
Abstract
This paper considers the equilibrium positions of particles in one dimension. Two forces act on the particles; a nonlocal repulsive particle-interaction force and an external force which pushes them to an impenetrable barrier. While the continuum limit as is known for a certain class of potentials, numerical simulations show that a discrete boundary layer appears at the impenetrable barrier, i.e. the positions of particles do not fit to the particle density predicted by the continuum limit. In this paper we establish a first-order -convergence result which guarantees that these particles converge to a specific continuum boundary-layer profile.
keywords: discrete-to-continuum; boundary layers; -convergence; -development.
MSC: 74Q05, 74G10, 49J45, 82C22.
1 Introduction
This paper contributes to a recent trend in interacting particle systems which aims to find more detailed information on the particle positions at equilibrium than the information which the continuum limit provides. There are roughly two directions which are currently pursued; convergence rates and particle patterns on mesoscopic scales. The studies on convergence rates (see, e.g. [BO20, EOS16, HvMP20, KvM21, PZ20, TS19, vM18]) aim to find a topology in which the distance between the configuration of particles and the continuum particle density can be measured and bounded by a small value which vanishes as . The studies on particle patterns zoom in on a mesoscopic scale, and reveal how the particles are distributed on this scale, either in the bulk (see the paper series started in [PS17, SS15]) or near the end of the support [GvMPS16, HCO10, HHvM18, Hud13]. This paper contributes to the latter, in which case we call the particle pattern a boundary layer. More precisely, this paper fills the important gap that was left open in [GvMPS16] on the characterization of boundary layers.
The gap in [GvMPS16]
To describe the gap in [GvMPS16], we first recall the corresponding setting. Consider many particles () confined to the half-line
We label their positions as and assume that they are ordered, i.e. , where
The discrete (i.e. ) particle interaction energy is given by
(1) |
where is an interaction potential, is a confining potential and is a parameter. The double sum accounts for each pair of two particles. Figure 1 illustrates typical examples for and . The assumptions and properties of and are roughly as follows. with and as . is normalized to , nonnegative, even and singular at with the bound
(2) |
for some , where is a parameter which bounds the strength of the singularity. We also assume that is non-increasing and convex. We state the precise list of assumptions and resulting properties in Section 3. With respect to [GvMPS16], we put more assumptions on , but allow for a general confining potential instead of the specific choice . This generalization of does not result in further complications. Instead, it clarifies the dependence of on the boundary layer.
One particular choice of which we have in mind is
(3) |
For this potential, is a model for the pile-up of dislocation walls at a lock. We refer to [GvMPS16] for the discussion of this model in the literature and its physical relevance. We show in Section 3 that it satisfies all the assumptions that we put on .
The asymptotic behaviour of the parameter in (1) as plays a decisive role for the limiting energy of as . This can be expected from (1) by noting that the scaled potential
(4) |
which has unit integral for all , is squeezed to a delta-peak at as . Hence, as increases, the particle interactions become more localized. In [GPPS13] the -limit of is obtained as . Depending on the asymptotic behavior of , five different limiting energies are obtained: two of these belong to the critical scaling regimes and as , and the other three belong to the three regimes separated by these two critical regimes (the outer two regimes require a rescaling of and ; we refer to [GPPS13] for the details).
In this paper we focus on the regime in between the two critical ones, i.e.
(5) |
In this regime the -limit of (see [GPPS13, Thm. 7]) is given by
(6) |
where is the space of probability measures on . The -norm is extended to measures (see (27) for details) and may be infinite. It is well-known (see, e.g. [KS80, Thm. 2.1] with minor modications to account for ) that the minimization problem of over has a unique minimizer and that its density is characterized by
(7) |
where the constant is such that . Obviously,
(8) |
Figure 2 illustrates .
As pointed out in [GvMPS16], the -convergence of is not completely satisfactory, because it does not detect any particle patterns on mesoscopic scales. For the scaling regime in (5), the numerical computations of the minimizer in [GvMPS16] indicate that particles are not distributed according to ; see Figure 3. The main result [GvMPS16, Thm. 1.1] captures the continuum boundary-layer profile according to which these particles are distributed. This profile is obtained by firstly proving a first-order -convergence result for the continuous counterpart of (see (9) below) and by secondly minimizing the first-order -limit. While Figure 3 suggests strongly that the obtained boundary-layer profile accurately describes the discrete boundary layer, any proof for this observation was left open. This is the gap in [GvMPS16] which we aim to fill in this paper.

The first-order -convergence result of [GvMPS16]
In order to describe this paper’s first-order -convergence result which will fill the gap in [GvMPS16], we first recall that of [GvMPS16]. By [GPPS13, Thm. 5] the -limit of in the regime is given by
(9) |
where is defined in (4). The energy is the continuous counterpart of which is considered in [GvMPS16]. It -converges to as (see [GvMPS16, Thm. 2.1]). In particular, this means that there exists a sequence such that
The idea of the authors of [GvMPS16] to upgrade this to a first-order -convergence result was to characterize the -term. They predicted from a priori computations that this term is , and that it is easier to replace by . This motivated them to consider the functional
They call the boundary-layer energy. The first-order -convergence of is simply the (zeroth-order) -convergence of .
For the -convergence of the topology needed to be chosen carefully. From the formal asymptotics in [Hal11] and their own numerical simulations the authors guessed that the width of the boundary layer is . This motivated them to use the following spatial rescaling. For a measure , let
(10) |
The inverse scaling is given by
Note that if has a density , then the density of satisfies
Using this scaling, the authors of [GvMPS16] employed the following change of variables:
By subtracting the bulk behaviour gets separated from the boundary layer.
For the signed Radon measures with total variation that growths linearly with , the authors used the vague topology. This topology is defined as follows on the space of signed Radon measures on . A sequence converges to vaguely (denoted by ) as if
The main result [GvMPS16, Thm. 1.1] states that -converges with respect to the vague topology to a certain limiting boundary-layer energy . This functional is defined on
(11) |
where are respectively the positive and negative part of such that . While may have infinite total variation, we have that and that the local bound on is translation invariant. For ,
(12) |
It is not obvious to extend this definition to , because need not be of finite total variation. We recall this extension briefly in Section 5.1. Finally, [GvMPS16] noted that has a unique minimizer (existence follows from the -convergence result in [GvMPS16] and uniqueness follows from the convexity of and the strict convexity of ), and that the continuous boundary-layer profile is given by
(13) |
Figure 3 and all other numerical simulations performed in [GvMPS16] suggest that gives a very good prediction for both the bulk and the boundary layer in the minimizer .
However, the match between and has only been observed and has not been proven. Hence, there is no guarantee that such a match extrapolates to any other choices for the potentials and and for the parameter . This motivates our aim to establish a first-order -convergence result for the discrete energy instead of its continuous counterpart .
First-order -convergence result of
To establish a first-order -convergence result for , we follow largely the same setup as the one just described. In fact, from Figure 3 we expect the same limiting boundary-layer energy . Also, there is a close connection between and , which can be seen as follows. Given , consider the corresponding empirical measure
(14) |
Then, we can express in terms of as
(15) |
where the diagonal
is removed from the integration domain to avoid self-interactions. Apart from removing the diagonal, the expressions for and are the same. Yet, the removal of the diagonal and the difference in the admissible sets on which and are defined are crucial. Indeed, concentrates around the diagonal as and thus careful analysis is required.
Following the procedure from [GvMPS16], we consider the blown-up energy difference and employ the following change of variables. For as in (14), we set
(16) |
Then, the discrete boundary-layer energy is defined on the admissible set
(17) |
and given by
(18) |
Note that if is constructed from by (16), then
(19) |
The main result of this paper in the following -convergence result of :
Theorem 1.1.
More precisely, the assumption on is equivalent to
The proof of Theorem 1.1 is given in Section 6 with preliminaries in Sections 4 and 5. It follows the proof in [GvMPS16] with major modifications to allow for the discreteness. Here, we briefly describe the main features of Theorem 1.1 and focus in particular on these major modifications.
First, we recall from [GvMPS16] that the expression for in (12) arises naturally when the right-hand side in (18) is explicitly expressed in terms of . In Section 4 we redo this computation, which in our case deals with the discrete setting and with a general confining potential .
The main difficulty with respect to [GvMPS16] is that the diagonal is removed from the integration domain (see (15)) and that the domain of is discrete (i.e. the degrees of freedom are empirical measures). To deal with this, we use essentially the particular regularization of constructed in [KvM21], which approximates from below as . Using this regularization, we add and subtract the contribution of the diagonal. By adding the diagonal, we can apply similar arguments as those in [GvMPS16] to establish the liminf inequality. However, needs to be chosen carefully. If is too small, then the contribution of the diagonal is too large and may not vanish in the limit. On the other hand, if is too large, then we cannot control the error made by the replacement of by . Balancing out these two errors results in the asymptotic upper bound on in Theorem 1.1. This bound is a stronger requirement than in (5), which is sufficient for the (zeroth-order) -convergence of to .
Establishing the limsup inequality is also significantly more challenging than in [GvMPS16]. The discreteness of forces us to discretize , which was not necessary in the continuous setting in [GvMPS16]. Since we blow up the energy difference by the factor , we need to show that the discretization error is asymptotically smaller than . This is much more intricate than for the zeroth-order -limit of (see [GPPS13, Thm. 7]), where it was sufficient to show that the discretization error simply vanishes as .
Discussion
In conclusion, Theorem 1.1 extends its continuous counterpart [GvMPS16, Thm. 1.1] (i.e. the -convergence of to ) in two manners. First, on a minor note, it allows for a general confining potential . This highlights the fact that the dependence of on is restricted to the single value , which depends nonlocally on (see (8) and Figure 2). Second, on a major note, Theorem 1.1 considers the discrete energy . As a consequence of Theorem 1.1, any sequence of minimizers of converges to . Since this convergence happens on the mesoscopic scale of the boundary layer, this proves that (see (13)) indeed describes the boundary layer which appears in . This gives the first theoretical motivation for the observations in Figure 3 and any other numerical computation in [GvMPS16] that fits to the regime of assumed in Theorem 1.1. This fills the main gap that was left open in [GvMPS16].
Yet, the story is not complete; [GvMPS16] contains a number of conjectures sparked by numerical simulations to which Theorem 1.1 does not provide an answer. Here, we focus on the main limitation of Theorem 1.1, which is the upper bound on . Indeed, the numerical simulations in [GvMPS16] suggest that is the correct boundary-layer profile for the whole regime of in (5). However, [GvMPS16, Table 1] suggests that the anticipated scaling of the energy difference, i.e. , ceases to hold at the upper bound on in Theorem 1.1. Hence, this upper bound is not simply an artefact of our proof. Looking deeper into the proof in Section 6, it seems that this upper bound is caused by the contribution to from a narrow region around the diagonal in the double integral in (15). A more precise treatment of this diagonal region could perhaps reveal a contribution of the right-hand side in (19) which diverges to as . Specifying this contribution, subtracting it from and proving -convergence of the resulting energy functional (provided that his is possible) would reveal that remains the correct boundary layer profile beyond the upper bound on in Theorem 1.1. Pursuing this direction is beyond our scope.
Organization of the paper
In Section 2 we set the notation. In Section 3 we state the precise assumptions on the potentials and , and derive further properties that follow from these assumptions. In Section 4 we rewrite defined in (18) explicitly in terms of , which will clarify the connection with the expression for in (12). In Section 5 we build the functional setting on which our proof of Theorem 1.1 relies. We also provide several a priori estimates. Finally, Section 6 is devoted to the proof of Theorem 1.1.
2 Notation
Here we list some symbols and abbreviations that we use throughout the paper.
smallest constant such that | As. 3.2(ii) | |
, | admissible sets for and | (17), (11) |
regularization parameter for and | (36) | |
modelling parameter | Thm. 1.1 | |
, | transforms of by scaling space by | (10) |
discrete energy | (1) | |
-limit of for | (9) | |
-limit of for | (6) | |
, | Fourier transform of ; | |
inverse Fourier transform of ; | ||
discrete boundary-layer energy | (18), (24) | |
continuum boundary-layer energy | (12), (30) | |
the Lebesgue measure on ; | ||
signed Radon measures on with support in | ||
positive and negative part of a | ||
measure ; | ||
is the set of probability measures | ||
minimizer of | (8) | |
rescaled version; | ||
‘convolutional square root’ of ; | (28), Lem. 5.1 | |
Hilbert space; | (32) | |
-norm; . |
We reserve for generic constants which do not depend on any of the relevant variables. We use in upper bounds (and think of it as possibly large) and in lower bounds (and think of it as possibly small). While may vary from line to line, in the same display they refer to the same value. If different constants appear in the same display, we denote them by .
To avoid clutter, we often omit the integration variable. For instance, we use
and extrapolate this notation to other integrands. Other than the framework of measures, we will also work with distributions. To connect the two notions, we often interpret measures on as distributions on supported in .
3 The potentials and
To the potential we add one more assumption to those mentioned in the introduction. We recall that is the constant in (7); see also Figure 2.
Assumption 3.1.
satisfies and as . Moreover, there exist finitely many disjoint closed intervals such that
(20) |
The assumption is not restrictive, as otherwise one can achieve this by adding a constant to . The assumption (20) is technical; it excludes pathological cases in which the graph of crosses the value infinitely many times. In fact, for the choice in [GvMPS16], (20) holds for . For as in Figure 2, (20) holds for .
In view of (8), Assumption 3.1 directly translates to assumptions on , independent of the assumptions on . Indeed, from (8) it is clear that Assumption 3.1 implies that is Lipschitz continuous, and that is uniformly continuous on . Hence, (20) implies that only at finitely many points is not of class .
Next we turn to the potential :
Assumption 3.2.
satisfies
-
(i)
(Evenness). is even;
-
(ii)
(Singularity). as , and there exist and such that for all
-
(iii)
(Convexity). is convex on and -convex near , i.e.
-
(iv)
(Integrability). is normalized to and has bounded first moment, i.e.
-
(v)
(Regularity). and .
First, we mention several properties of which follow from Assumption 3.2. The evenness, convexity and integrability imply that is non-increasing on and that is real-valued, nonnegative and even, which is sufficient for Assumption 3.2(v) to be well-defined. A less obvious consequence is Lemma 3.3.
Lemma 3.3.
There exists a constant such that for all
Proof.
Since is even and real-valued, it is enough to focus on . [KvM20, (A.3)] provides the characterization
Since the integrand is nonnegative, we may bound it from below by shrinking the integration domain. Then, on , we bound . For we note from Assumption 3.2(iii) that
where equals if the statement is true and otherwise. Then,
We split two cases. If , then only the term corresponding to is nonzero, and the right-hand side equals
If , then we estimate
∎
Next we compare Assumption 3.2 to the assumptions on made in [GvMPS16], which are weaker. Indeed, in [GvMPS16] Assumption 3.2(ii) and the regularity on are not required, and Assumption 3.2(iii) is relaxed to the requirement that is non-increasing. While [GvMPS16] has a further assumption that can be approximated from below by a certain class of functions, we show that this holds under Assumption 3.2 by constructing such an approximation explicitly; see (36).
Next we motivate the assumptions which are new with respect to [GvMPS16]. We believe that these additional assumptions are minor, and still allow for most of the potentials in practice which satisfy the assumptions in [GvMPS16]. Regarding Assumption 3.2(ii), it is obvious from the bound on in Theorem 1.1 that a bound on the singularity of is needed. The requirement as might not be necessary. However, including this case in the proof would require a further case splitting. Since we are not aware of any application for this case, we omit it. While the convexity in Assumption 3.2(iii) is new, it captures the following three assumptions in [GvMPS16]: (see the proof of Lemma 3.3), is non-increasing, and can be approximated from below by a special class of functions. The local -convexity is a technical addition which simplifies several steps in the proof of Theorem 1.1. Finally, in Assumption 3.2(v), the regularity on is only a small upgrade of , which follows from convexity. The regularity of is required in [GvMPS16] to extend from to . We further exploit this assumption when proving properties of the regularization . This is the single assumption which can be hard to check in practice.
Finally, we show that defined in (3) satisfies Assumption 3.2. We recall from [GPPS13] that is strictly convex, has a logarithmic singularity (in particular, ) and exponential tails. Then, the only non-trivial property left to check is the regularity of with . By the strict convexity and the exponential tails of , it follows that is positive, and thus . To extend this to large , we recall from [GPPS13, App. A.1] that
Note that and as . Then, we obtain from
4 Explicit expression of
Here we derive an explicit expression for in terms of and motivate the prefactor in (18). By scaling back, note from (16) that can be written as
for some empirical measure of the form (14). Then, in view of the right-hand side in (19), we set and compute
where we recall that and are extended from to by .
Next we rewrite the second term. With this aim, we set
(21) |
and expand
(22) |
The first term equals
(23) |
For the integrand of the third term in (22), we note from (8) that
Collecting our computations, we obtain
Multiplying by and changing variables (recall ), we get
(24) |
For later use we note that the integral in the second term can be rewritten as (recall (23))
(25) |
Note that the first two terms in (24) resemble the expression of in (30). This motivates the scaling of the energy difference in (18) by . We treat the latter three terms in (24) as error terms when proving Theorem 1.1. While the third term obviously vanishes as , the other two terms may not for certain sequences . We rely on the fact that the second term is nonnegative, and that the integrand in the first term is expected to be small because as . We give a precise bound later in Lemma 5.6.
5 Functional setting and preliminaries
In Section 5.1 we recall from [GvMPS16] the necessary functional framework to extend the definition of in (12) to . Since this functional framework also facilitates the statements and proofs of several preliminary estimates, we treat them in the subsequent Section 5.2. In this functional framework we identify measures on as tempered distributions on supported in .
5.1 Proper definition of
Since is the same as in [GvMPS16], we briefly recall the extension of the definition in (12) on to . Ideally, if there exists a function such that , then for we have for the interaction term that
(26) |
This expression can be extended to distributions by noting that
(27) |
However, from Assumption 3.2 it is not clear whether such a function exists.
One way to avoid characterizing is to work in Fourier space. Since convolution transforms into multiplication by the Fourier transform, the linear operation of convolving by turns into multiplication by , which is a function due to Assumption 3.2(v). Precisely, we set
(28) |
where . Then, by translation to Fourier space, we observe that (26) turns into
(29) |
Together with the observation in (25) this yields
(30) |
for all , where
(31) |
To extend to , we show that the linear term in (30) is bounded and that the operator can be extended to . We do this in Lemmas 5.1 and 5.2. For later use, we state these lemmas in a general form. With this aim, we introduce the Hilbert spaces
for all . In case the functions are real-valued, we set
(32) |
Note that for all , where is the space of Schwarz functions. Note from the Fourier transform property
that is an invertible bounded linear operator from to . We further set as the dual of with respect to the -topology, and as the space of all bounded linear operators from to .
Lemma 5.1 ([GvMPS16, Lem. A.2]).
The operator and the set satisfy
-
(i)
;
-
(ii)
For each and each , is symmetric, and can be extended to by
Lemma 5.2 ([GvMPS16, Lem. 3.4]).
There exist constants with
such that for all and all it holds that
(33) | ||||
(34) |
5.2 A priori bounds on
First we state the counterpart of Lemma 5.1(i) in the discrete setting. Since Dirac-delta measures are included in , we obtain
(35) |
While , it is pointless to consider on , because for each due to the discreteness and the singularity of at . This is the crucial difference with the continuous setting in [GvMPS16] in which is a small perturbation of .
To deal with the discreteness, we construct a careful regularization of . With this aim, let be as in Assumption 3.2(iii). For we define the regularization
(36) |
and consider the even extension of to . Figure 4 illustrates . Note that on , is a parabola which is tangent to at . We also set
Lemma 5.3 lists several properties of .
Lemma 5.3 (Properties of ).
There exist constants such that for all small enough
-
(i)
(Pointwise bounds). and ;
- (ii)
-
(iii)
(Narrow support). ;
-
(iv)
(Fourier transform). is real-valued and even, and satisfies
-
(v)
(, bounds). and .
Proof.
Properties (i), (ii), (iii) and the fact that is real-valued and even hold by construction. From these properties, we observe that the proof of Lemma 3.3 also applies to ; this proves the lower bound in (iv). The upper bound follows simply from and (i). The bound on in (v) follows from (iii) through
To prove the bound on in (v), we observe from (36) that
By the Mean Value Theorem and the fact that is non-increasing, we get
This completes the proof of Lemma 5.3. ∎
Lemma 5.3(iv) allows us to define, similarly to , the convolutional square root operator
where . As in (29), it is easy to see (in Fourier space) that and
(37) |
Lemma 5.4 states further properties of .
Lemma 5.4.
The operator defined above satisfies
-
(i)
For each and all small enough, is symmetric, and can be extended to as in Lemma 5.1;
-
(ii)
for each ;
-
(iii)
For all small enough, each and all , there holds ,
-
(iv)
There exists such that for all Schwarz functions there exists such that for all
Proof.
Since is even and bounded, (i) and (ii) follow from the same argument used in the proof of [GvMPS16, Lem. A.2(ii)]. In particular,
Next we proof (iii). It is clear that holds in distributional sense. Note that (37) defines a seminorm, which can be extended to distributions similarly to (27). Using this and applying (i), we write
where . We claim that is dense in . From this claim, (iii) follows by applying (27). To prove the claim, we note that, by translating it to Fourier space, it is equivalent to the claim that is dense in . This is easily seen to be true; for , set and note from the lower bound in Lemma 5.3(iv) that for all .
Finally we prove (iv). Note that
(38) |
Recalling , we set and compute
where . We observe from Lemma 5.3(i),(iii) that for any
This will eventually result in the prefactor in (iv).
Next we bound . We recall from Assumption 3.2(v) that
and from Lemmas 3.3 and 5.3(iv) that
Here and henceforth, we often abuse notation by removing the variable from the notation for functions in Fourier space. Then,
where is independent of and
Using this, we obtain for the -part of the norm in (38) that
For the -part of the norm, we compute
(39) |
Writing , all terms in the expression for can be bounded from the estimates above. This yields
Hence,
Finally we bound the -part of the norm. By the estimates obtained so far,
(40) |
In preparation for estimating the second derivative, we compute (using the expression in (39))
Then, rewriting
we estimate all terms in the expression for by the bounds obtained above. This yields
Returning to (40), we obtain
Plugging this estimate into (40) completes the proof of (iv). ∎
This completes the preliminaries on the regularization . Next we apply them to construct tools for the proof of Theorem 1.1. The first of these tools is a crucial estimate in the proof of compactness. It is the discrete counterpart of [GvMPS16, Lem. 3.3].
Lemma 5.5.
There exists a constant such that for all small enough and all
Proof.
The proof is a modification of the proof of [GvMPS16, Lem. 3.3]. Let be arbitrary. Using that is non-increasing on , we obtain
Taking small enough such that , Lemma 5.4(iii) implies
Then, applying the Cauchy-Schwarz Inequality and Lemma 5.4(ii)
(41) |
which shows the desired estimate for the first term in the display in Lemma 5.5.
The local bound on in Lemma 5.5 turns out useful when passing to the limit in the second and third term in the expression for in (24). This is made precise in the following two lemmas.
Lemma 5.6.
Proof.
Let . Since is Lipschitz continuous, we obtain
(42) |
which is bounded due to Assumption 3.2(iv). Hence,
(43) |
To prove the convergence statement in Lemma 5.6, we improve the bound on in (42). Let be arbitrary. By (8) and (20) there exist finitely many points such that is uniformly continuous on . Let where . Then, for all ,
where the bounded function vanishes as uniformly in . Using this, we estimate, similarly to (42),
Using this, we sharpen the bound in (43) by
The second term vanishes as . For the first term, we note that
Covering each interval with many intervals of length , we use the given bound on to continue this estimate by
In conclusion,
Since is arbitrary, Lemma 5.6 follows. ∎
Lemma 5.7.
Let be such that as . If
then (recall )
Proof.
We follow the proof in [GvMPS16] for the continuum setting and present it here in more detail. Take a continuous cut-off function
(44) |
We recall from (25) that
(45) |
where is an arbitrary constant. Since , we obtain from that
For the second term in (45), we use (33) to estimate
where as . By a similar argument, it follows from that
Tracing these observations back to (45), we conclude
Since is arbitrary, Lemma 5.7 follows. ∎
6 Proof of Theorem 1.1
This section is devoted to the proof of Theorem 1.1. Theorem 1.1 consists of three statements: compactness, the liminf inequality and the limsup inequality. We prove these three statements respectively in Sections 6.1, 6.2 and 6.3 for the power-law case (see Assumption 3.2(ii)). In Section 6.4 we show that with minor modifications the proof for the logarithmic case follows.
6.1 Compactness
Let be such that . We start from the expression for in (24). Since the fourth and fifth term in (24) are nonnegative (recall (7)), we may neglect them. By Lemma 5.6 the third term is uniformly bounded. Hence, it is sufficient to focus on the first two terms in (24), which we label .
Recalling the regularization defined in (36), we expand
(46) |
For the second term, note from Lemma 5.4(iii) that
The first and third terms in (46) can be bounded from below by small constants. Indeed, using , we bound the third term by
For the first term in (46), we recall and expand :
Since , the first and third term are nonnegative. Using , we estimate the second term by
Hence, the first and third term in (46) are bounded from below by
We choose such that this quantity is maximal. Up to a constant, this yields
Then, by the assumption on in Theorem 1.1, we obtain
Collecting these estimates, we obtain from (46) that
(47) |
For the fourth term in (47), we use a rougher estimate. By Lemma 5.5,
for all large enough. Then, together with the first two terms in (47), we obtain
(48) |
By (48), is bounded. This implies two useful properties. First,
in as along a subsequence (not relabelled) for some . Then, by Lemma 5.5, is bounded for any . Hence,
along a further subsequence as .
It is left to show that , i.e.
(49) |
Since and is continuous, it follows that
as . Then, together with , we obtain
as . In particular, , which shows the first statement in (49). To prove the second statement, we take arbitrary, and take a test function which satisfies and
Then,
which by Lemma 5.5 and (48) is bounded uniformly in . This implies the second statement in (49).
6.2 Liminf inequality
To prove the liminf inequality in Theorem 1.1, let be given. We may assume that is bounded along a subsequence in (not relabelled), as otherwise the liminf inequality is trivial. Then, as in the compactness proof, we obtain (48), which shows that is bounded in . Then, Lemma 5.5 implies
(50) |
Let be as in (46). First, we observe that
(51) |
Indeed, in the compactness proof we already showed that the fourth and fifth term in (24) are nonnegative. By (50) and Lemma 5.6, the third term vanishes as .
Next we bound the right-hand side in (51) from below. As in the proof for the compactness, we obtain (47). Together with Lemma 5.7 this yields
In particular, in as along a subsequence to some . It is left to prove that . With this aim, let be a test function. Using (35) and Lemma 5.4(i) we obtain
(52) |
where is the cut-off function introduced in (44), and is an arbitrary constant. We pass to the limit in both terms separately.
For the first term in (52), we note from Lemma 5.4(iv) that in as . Together with this yields
For the second term in (52), we set and obtain from (34), Lemma 5.5, and the triangle inequality that
(53) |
For the second term, we set and compute
Since , this value vanishes as . For the first term in (53), we obtain from Lemma 5.4(iv) that
which vanishes as uniformly in . In conclusion, by tracing these observations back to (52), we obtain
(54) |
where as .
6.3 Limsup inequality
We first assume and treat the special case afterwards. Since is the same as in [GvMPS16] and only depends on through the constant , we may use the density result in [GvMPS16]. This result states that it is sufficient to prove the limsup inequality only for those for which , for some and on for some . In particular, we treat as a density. Since is continuous as a functional on , we may further assume that .
We first treat the case , and comment on the case afterwards. To choose , we note from (17) that can be parametrized by through
(55) |
where we recall that depends on . We choose such that
(56) |
Note from that such an exists. By taking large enough, we may further assume that and that
(57) |
Next we prove several properties of and . We observe from (56) that
(58) |
Hence,
(59) |
To prove
(60) |
we take a test function and set . Taking large enough such that on and on , we note from (56) that
(61) |
which vanishes as . Then, from
we obtain by the continuity of that the right-hand side vanishes as . This proves (60).
Next we prove the limsup inequality in Theorem 1.1 for as constructed above. With this aim, we treat all five terms of in (24) separately. The latter four terms all converge as . Indeed, the fifth term in (24) vanishes as . By (57) the fourth term equals . Due to (59), Lemma 5.6 implies that the third term vanishes as . Due to (59) and (60), Lemma 5.7 guarantees the convergence of the second term.
Therefore, it is sufficient to prove the limsup inequality only for the first term in (24). We first show that for all small enough
(62) |
Let be the smallest integer for which . Using that is even, we expand
(63a) | ||||
(63b) | ||||
(63c) | ||||
(63d) | ||||
(63e) | ||||
(63f) |
The second term in (63a) is negative; we simply bound it from above by . By (60) and , we obtain that as . By [AFP00, Thm. 1.59] we then also have that as . Since is continuous, this implies for the first term in (63a) that
which equals the integral in the right-hand side of (62).
Next we bound (63b). Neglecting the negative cross terms, we estimate
(64) |
For the second term, we recall from Lemma 5.3(v) that . Since for large enough, we obtain
The first term in (64) is the discrete counterpart of the second term, and can be treated similarly. Relying on (58) and , we bound it by
Hence, (63b) is bounded by uniformly in . In view of (62), it is therefore left to show that the limsup of the remaining terms in (63) are nonpositive.
We start with (63d). Expanding and using that is even, we rewrite
(65) |
Using (56) and the fact that is decreasing, we obtain for the integrals inside the parentheses that
(66) |
and, similarly,
for all . Hence, . For , we observe from and (56) that
(67) |
Hence, , which by the assumption on in Theorem 1.1 vanishes as . In conclusion, the limsup of (63d) is nonpositive.
For (63f), we note that is disjoint with . Hence,
(68) |
Then, since
(69) |
we get . Since is Lipschitz continuous, this implies that
Then, similarly to (67), we estimate the right-hand side of (68) as
which vanishes as .
Next we treat (63c). We split in the inner integral. For the first part, we expand as in (65),
A similar argument as that in (66) yields and
Noting from (58) that
we obtain .
The second part of (63c) equals (setting )
(70) |
To estimate this in absolute value, we split the sum in two parts. Let and take large enough such that (recall that ) and such that (61) holds for all . Then,
and, recalling (58),
Hence, the limsup of the second part of (63c) is nonpositive. We conclude that the limsup of (63c) is nonpositive.
Finally, for (63e), we observe that the inner integral
is nonpositive and non-increasing on . Then, using a similar argument as for (63c) (simplifications are possible), we conclude that the limsup of (63e) is nonpositive. This completes the proof of (62).
In conclusion, we have proved that
for all small enough. Since and in as , we conclude the limsup inequality in Theorem 1.1 by taking and applying (29).
It is left to treat the cases and . Since implies , these two cases are mutually exclusive. We start with the case . We follow the proof for the case with minor modifications. By density, instead of assuming , we may assume that on , where is a monotone cut-off functions which equals on and on . Under this assumption, (61) does not hold for large , and thus we need to construct a different argument for (60) and for showing that the limsup of (63c) is nonpositive.
To prove (60), we take a test function and set . From the proof of (60) we observe that on implies that as . For the integral over , we note from the Lipschitz continuity of and that
Then, by (56),
Hence,
We conclude (60).
Next we show that the limsup of (63c) is nonpositive. We can follow the proof for the case up to the term in (70). Since implies , we observe that is nonnegative and non-increasing on . Hence, by a similar argument as that in (66), it follows that . This concludes the proof of the limsup inequality in Theorem 1.1 in the case .
It is left to prove the limsup inequality for the case . We largely follow the proof for the case , and focus on the modifications. For the choice of , we define as in (56), where is the smallest integer at which
For we set
(71) |
and take as in (55). Clearly, properties (58), (59) and (60) are still satisfied. To obtain the discrete equivalent of (69), we compute
Hence,
(72) |
Finally, instead of (57), we now have
(73) |
Similarly to the previous case, we pass to the limit in all five terms in (24) separately. By Lemmas 5.6 and 5.7 we obtain the same limit for the second, third and fifth term. Using (73), the fourth term equals
To bound the sum in the right-hand side, note from that
Since by (72)
it follows from that the summand vanishes as . Hence, the fourth term in (24) also vanishes as .
We treat the first term in (24) similarly as in (62) and (63). The modifications to (63) are as follows. First, we replace by in all integration domains. Then, we replace in (63e) and (63f) by in the integrals over . Finally, we add two new terms (see (74)) to account for . With these modifications, (63a)–(63f) can be treated analogously. The two new terms that need to be added are
(74) |
Using (71) and (72) we estimate the first term by
For the second term in (74), we apply a similar argument as for (63c). This yields
and
Hence, (74) is bounded from above by
This completes the proof of the limsup inequality in Theorem 1.1.
6.4 The case
Here we prove Theorem 1.1 for the case . The proof is the same as for the case , except for minor computational modifications. All these modifications are ramifications from the difference in the bound on in Assumption 3.2(ii), which in the current case produces logarithms. The ramifications in the preliminary estimates are the statement of Lemma 5.3(v), which changes into
(75) |
and the statement of Lemma 5.4(iv). By observing that , it follows from the proof of Lemma 5.4(iv) that the corresponding statement becomes
(76) |
In the compactness proof, by taking again , we observe from (75) that the term becomes
Applying the asymptotic bound on in Theorem 1.1, we obtain
which is sufficient for continuing the argument in the proofs for the compactness and the liminf inequality.
For the liminf inequality, we only need that (76) vanishes as , which is obvious. For the limsup inequality, the bounds in (75) yield that (63b) is bounded by . Hence, we may replace the term in (62) by . Also, the bound in (67) changes by (75) into
From this estimate, we obtain for the term in (65) that
which is sufficient for the proof of the limsup inequality.
Acknowledgements
The author gratefully acknowledges support from JSPS KAKENHI Grant Number JP20K14358.
References
- [AFP00] L. Ambrosio, N. Fusco, and D. Pallara. Functions of Bounded Variation and Free Discontinuity Problems. Clarendon Press, Oxford, 2000.
- [BO20] J. Braun and C. Ortner. Sharp uniform convergence rate of the supercell approximation of a crystalline defect. SIAM Journal on Numerical Analysis, 58, 2020.
- [EOS16] V. Ehrlacher, C. Ortner, and A. V. Shapeev. Analysis of boundary conditions for crystal defect atomistic simulations. Archive for Rational Mechanics and Analysis, 222(3):1217–1268, 2016.
- [GPPS13] M. G. D. Geers, R. H. J. Peerlings, M. A. Peletier, and L. Scardia. Asymptotic behaviour of a pile-up of infinite walls of edge dislocations. Archive for Rational Mechanics and Analysis, 209:495–539, 2013.
- [GvMPS16] A. Garroni, P. van Meurs, M. A. Peletier, and L. Scardia. Boundary-layer analysis of a pile-up of walls of edge dislocations at a lock. Mathematical Models and Methods in Applied Sciences, 26(14):2735–2768, 2016.
- [Hal11] C. L. Hall. Asymptotic analysis of a pile-up of regular edge dislocation walls. Materials Science and Engineering: A, 530:144–148, 2011.
- [HCO10] C. L. Hall, S. J. Chapman, and J. R. Ockendon. Asymptotic analysis of a system of algebraic equations arising in dislocation theory. SIAM Journal on Applied Mathematics, 70(7):2729–2749, 2010.
- [HHvM18] C. L. Hall, T. Hudson, and P. van Meurs. Asymptotic analysis of boundary layers in a repulsive particle system. Acta Applicandae Mathematicae, 153(1):1–54, 2018.
- [Hud13] T. Hudson. Gamma-expansion for a 1D confined Lennard-Jones model with point defect. Networks & Heterogeneous Media, 8(2):501–527, 2013.
- [HvMP20] T. Hudson, P. van Meurs, and M. A. Peletier. Atomistic origins of continuum dislocation dynamics. ArXiv: 2001.06120, 2020.
- [KS80] D. Kinderlehrer and G. Stampacchia. An introduction to variational inequalities and their applications. Academic Press. NY, London, 1980.
- [KvM20] M. Kimura and P. van Meurs. Regularity of the minimiser of one-dimensional interaction energies. ESAIM: Control, Optimisation and Calculus of Variations, 26:27, 2020.
- [KvM21] M. Kimura and P. van Meurs. Quantitative estimate of the continuum approximations of interacting particle systems in one dimension. SIAM Journal on Mathematical Analysis, 53(1):681–709, 2021.
- [PS17] M. Petrache and S. Serfaty. Next order asymptotics and renormalized energy for Riesz interactions. Journal of the Institute of Mathematics of Jussieu, 16(3):501–569, 2017.
- [PZ20] L. Pronzato and A. Zhigljavsky. Bayesian quadrature, energy minimization, and space-filling design. SIAM/ASA Journal on Uncertainty Quantification, 8(3):959–1011, 2020.
- [SS15] E. Sandier and S. Serfaty. 1D log gases and the renormalized energy: crystallization at vanishing temperature. Probability Theory and Related Fields, 162(3-4):795–846, 2015.
- [TS19] K. Tanaka and M. Sugihara. Design of accurate formulas for approximating functions in weighted Hardy spaces by discrete energy minimization. IMA Journal of Numerical Analysis, 39(4):1957–1984, 2019.
- [vM18] P. van Meurs. Convergence rates for discrete-to-continuum limits in 1D particle systems. In Mathematical Analysis of Continuum Mechanics and Industrial Applications II: Proceedings of the International Conference CoMFoS16, pages 181–193. Springer, 2018.