On a Class of Nonlocal Obstacle Type Problems Related to the Distributional Riesz Fractional

Catharine W.K. Lo∗ and Jos´eFrancisco Rodrigues†

Abstract In this work, we consider the nonlocal obstacle problem with a given obstacle ψ in a bounded Lipschitz d s s domain Ω in R , such that Kψ = {v ∈ H0 (Ω) : v ≥ ψ a.e. in Ω} 6= ∅, given by

s s u ∈ Kψ : hLau, v − ui ≥ hF, v − ui ∀v ∈ Kψ,

−s s for F in H (Ω), the dual space of the fractional H0 (Ω), 0 < s < 1. The nonlocal s −s operator La : H0 (Ω) → H (Ω) is defined with a measurable, bounded, strictly positive singular kernel d d a(x, y) : R × R → [0, ∞), by the

hLau, vi = P.V. v˜(x)(˜u(x) − u˜(y))a(x, y) dy dx = Ea(u, v), ˆ d ˆ d R R which is a (not necessarily symmetric) Dirichlet form, whereu, ˜ v˜ are the zero extensions of u and v ˜ s s s outside Ω respectively. Furthermore, we show that the fractional operator LA = −D · AD : H0 (Ω) → H−s(Ω) defined with the distributional Riesz fractional Ds and with a measurable, bounded matrix A(x)

corresponds to a nonlocal operator LkA with a well defined integral singular kernel a = kA. The corresponding s-fractional obstacle problem for L˜A is shown to converge as s % 1 to the obstacle problem 1 in H0 (Ω) with the operator −D · AD given with the classical gradient D. We mainly consider obstacle-type problems involving the bilinear form Ea with one or two obstacles, as well as the N-membranes problem, thereby deriving several results, such as the weak maximum prin- ciple, comparison properties, approximation by bounded penalization, and also the Lewy-Stampacchia inequalities. This provides regularity of the solutions, including a global estimate in L∞(Ω), local H¨older 2s,p regularity of the solutions when a is symmetric, and local regularity in fractional Sobolev spaces Wloc (Ω) 1 s and in C (Ω) when La = (−∆) corresponds to fractional s-Laplacian obstacle-type problems

s s s s 2 d d u ∈ Kψ(Ω) : (D u − f) · D (v − u) dx ≥ 0 ∀v ∈ Kψ for f ∈ [L (R )] . ˆ d R These novel results are complemented with the extension of the Lewy-Stampacchia inequalities to the s arXiv:2101.06863v4 [math.AP] 15 Sep 2021 order dual of H0 (Ω) and some remarks on the associated s-capacity and the s-nonlocal obstacle problem for a general La.

1 Introduction

Fractional problems with obstacle-type constraints were first considered by Silvestre as part of his thesis in 2005 in [54], which was published in 2007 in [53]. Since then, many obstacle-type problems with various nonlocal operators have been considered, extensively for the fractional Laplacian (such as in [13], [18], [39], [40] and [46]), as well as for other nonlocal operators (see, for example, [1], [5], [12], [17], [48] and [49]),

∗CMAFcIO – Departamento de Matem´atica,Faculdade de Ciˆencias,Universidade de Lisboa P-1749-016 Lisboa, Portugal Email address: [email protected] †CMAFcIO – Departamento de Matem´atica,Faculdade de Ciˆencias,Universidade de Lisboa P-1749-016 Lisboa, Portugal Email address: [email protected]

1 which are mainly generalizations of the fractional Laplacian to nonlocal one-variable kernels K(y) satisfying homogeneous and symmetry properties (see in particular, [1] and [12]). Recently, in a series of two interesting papers [50] and [51], Shieh and Spector have considered a new class of fractional partial differential equations based on the distributional Riesz fractional . These fractional operators satisfy basic physical invariance requirements, as observed by Silhavy [52], who developed a fractional vector calculus for such operators. Instead of using the well-known fractional Laplacian, their starting concept is the distributional Riesz fractional gradient of order s ∈ (0, 1), which will be called here the s-gradient Ds, for brevity: for u ∈ Lp(Rd), p ∈ (1, ∞), we set s s ∂ u ∂ Dj u = s = (I1−s ∗ u), 0 < s < 1, j = 1, . . . , d, (1.1) ∂xj ∂xj

where ∂ is taken in the distributional sense, for every v ∈ C∞( d), ∂xj c R * + * + ∂su ∂v ∂v s , v = − (I1−s ∗ u), = − (I1−s ∗ u) dx, ∂x ∂x ˆ d ∂x j j R j with Is denoting the Riesz potential of order s, 0 < s < 1:   Γ d+s+1 u(y) d 2 s − 2 (Is ∗ u)(x) = cd,1−s d−s dy, with cd,s = 2 π 1−s  . (1.2) ˆ d |x − y| R Γ 2 ∞ Conversely, by Theorem 1.12 of [50], every u ∈ Cc (Ω) can be expressed as

d X ∂su u = I ∗ R , (1.3) s j ∂xs j=1 j where Rj is the Riesz transform, which we recall, is given by

 d+1  Γ 2 y R f(x) := lim j f(x − y) dy, j = 1, . . . , d. j (d+1)/2 d+1 π →0 ˆ{|y|>} |y| Thus, we can write the s-gradient (Ds) and the s-divergence (Ds·) for sufficiently regular functions u and vector φ ([16], [50], [51], [52]) in integral form, respectively, by

zu(x + z) u(x) − u(y) x − y s : D u(x) = cd,s lim d+s+1 χ(0, z) dz = cd,s d+s dy →0 ˆ d |z| ˆ d |x − y| |x − y| R R and z · φ(x + z) φ(x) − φ(y) x − y s : D · φ(x) = cd,s lim d+s+1 χ(0, z) dz = cd,s d+s · dy, →0 ˆ d |z| ˆ d |x − y| |x − y| R R where χ(x, z) is the characteristic function of the set {(x, z) : |z − x| > } for  > 0. As it was shown in s ∞ d [50], D has nice properties for u ∈ Cc (R ), namely it coincides with the fractional Laplacian as follows: (−∆)su = −Ds · Dsu,

where, for 0 < s < 1,

s 2 u(x) − u(y) 1 2 u(x + y) + u(x − y) − 2u(x) (−∆) u(x) = cd,s lim d+2s χ(x, y) dy = cd,s d+2s dy. →0 ˆ d |x − y| 2 ˆ d |y| R R Observe that for the s-gradient (Ds) and the s-divergence (Ds·) we need to consider the Cauchy principal value (P.V.) in the first expressions, but not in the second ones. This is because for the second expressions, we can estimate the integrand, as in [16], by separating the into the parts {|y − x| ≤ 1} and 1 −s {|y − x| > 1}. Then the first integral can be controlled by ωd Lip(u) 0 r dr, while the second integral ´ 2 +∞ −(1+s) can be controlled by 2ωdkuk ∞ d r dr, where Lip(u) is the Lipschitz constant for the function L (R ) 1 ´ u and ωd is the spherical measure ωd = {|x|=1} dσ. Therefore, the second expressions are well-defined for ´ ∞ d all Lipschitz functions u with compact support, in particular for u ∈ Cc (R ). Similarly, the fractional Laplacian for smooth u, by Lemma 3.2 of [23], has two representations, with the second one being Lebesgue integrable by using a second order Taylor expansion. As a matter of fact, the s-divergence, s-gradient and s-Laplacian are linear operators from C∞ functions with compact support into C∞ functions that are rapidly decreasing at ∞ and are in Lp(Rd) for any p ∈ [1, ∞]. ∞ d Observe that for u, v ∈ Cc (R ), by using the duality between s-divergence and s-gradient as in Lemma 2.5 of [16], we have

2 Dsu · Dsv = 2 v(−∆)su ˆ d ˆ d R R   2 u(y) − u(x) u(x) − u(y) = cd,s lim v(y) d+2s χ(x, y) dx dy + lim v(x) d+2s χ(x, y) dy dx ˆ d →0 ˆ d |x − y| ˆ d →0 ˆ d |x − y| R R R R   (1.4) 2 u(y) − u(x) u(x) − u(y) = cd,s lim v(y) d+2s χ(x, y) dx dy + lim v(x) d+2s χ(x, y) dx dy →0 ˆ d ˆ d |x − y| →0 ˆ d ˆ d |x − y| R R R R 2 (u(x) − u(y))(v(x) − v(y)) = cd,s d+2s dx dy, ˆ d ˆ d |x − y| R R where we have used the above definitions together with the Lebesgue and Fubini theorems. s In this work, we are concerned with the classical fractional Sobolev space H0 (Ω) in a bounded domain Ω ⊂ Rd with Lipschitz boundary, for 0 < s < 1, defined as

k·k s ∞ Hs H0 (Ω) := Cc (Ω) ,

with 2 2 s 2 kuk s =kuk 2 d +kD uk 2 d , H L (R ) L (R ) where u is extended by 0 in Rd\Ω, so that this extension is also in Hs(Rd). By the Sobolev-Poincar´e s inequality (see Theorem 1.7 of [50] and Lemma 3.2), we may consider the space H0 (Ω) with the following equivalent norms, owing to (1.4),

2 2 2 2 s 2 cd,s 2 cd,s (u(x) − u(y)) s : : kukH (Ω) =kD ukL2( d) = [u]s, d = d+2s dx dy. (1.5) 0 R 2 R 2 ˆ d ˆ d |x − y| R R On the other hand, since the Riesz kernel is an approximation to the identity as 1−s → 0, the s-derivatives approach the classical derivatives as s → 1 in appropriate spaces, as it was observed in Rodrigues-Santos [44], Comi-Stefani [15] and Belido et al [8]. s −s We can subsequently denote the dual space of H0 (Ω) by H (Ω) for 0 < s ≤ 1. Then, by the Sobolev- Poincar´einequalities, we have the embeddings

s q 2# −s s 0 H0 (Ω) ,→ L (Ω),L (Ω) ,→ H (Ω) = (H0 (Ω))

∗ ∗ 2d # 2d d ∗ for 1 ≤ q ≤ 2 , where 2 = d−2s and 2 = d+2s when s < 2 , and if d = 1, 2 = q for any finite q and # 0 q 1 ∗ # 1 2 = q = q−1 when s = 2 and 2 = ∞ and 2 = 1 when s > 2 . We recall that those embeddings are compact for 1 ≤ q < 2∗ (see for example, Theorem 4.54 of [21]). In the whole paper we use 2# to indicate this number that depends on d ≥ 1 and 0 < s ≤ 1. We consider the closed convex set

s s Kψ = {v ∈ H0 (Ω) : v ≥ ψ a.e. in Ω},

s with a given obstacle ψ, such that Kψ 6= ∅, and the obstacle problem

s s u ∈ Kψ : hLau, v − ui ≥ hF, v − ui ∀v ∈ Kψ, (1.6)

3 −s s −s for F in H (Ω). Here, the nonlocal operator La : H0 (Ω) → H (Ω) is a generalization of the fractional Laplacian for a measurable, bounded, strictly positive kernel a : Rd × Rd → [0, ∞) satisfying (2.2), and is s defined in the duality sense for u, v ∈ H0 (Ω), extended by zero outside Ω:

hLau, vi := lim v˜(x)(˜u(x) − u˜(y))a(x, y)χε(x, y) dy dx = P.V. v˜(x)(˜u(x) − u˜(y))a(x, y) dy dx. ε→0 ˆ d ˆ d ˆ d ˆ d R R R R (1.7) Physically, the operator La corresponds to the class of uniformly irreducible random walks that admit a cycle decomposition with bounded range, bounded length of cycles, and bounded jump rates [22]. To better characterize the properties of the operator La, in Section 2, we show that the bilinear form

Ea(u, v) := hLau, vi = P.V. v˜(x)(˜u(x) − u˜(y))a(x, y) dy dx ˆ d ˆ d R R s s is a (not necessarily symmetric) Dirichlet form over H0 (Ω) × H0 (Ω), whereu ˜ andv ˜ are the zero extensions of u and v outside Ω respectively. This provides us with many known properties of Dirichlet forms that can be applied to the bilinear form Ea, including the truncation property and some regularity results [29][37]. As s a corollary, we obtain that Ea is a closed, coercive, strictly T-monotone and regular Dirichlet form in H0 (Ω). Furthermore, we use the results of the nonlocal vector calculus developed by Du, Gunzburger, Lehoucq and ˜ s −s coworkers in [19], [24], [25] and [31] to show that the fractional operator LA : H0 (Ω) → H (Ω), defined by

˜ s s s hLAu, vi := A(x)D u · D v dx, ∀u, v ∈ H0 (Ω), (1.8) ˆ d R corresponds to a nonlocal integral operator LkA with a (not necessarily symmetric) singular kernel kA(x, y) defined by (2.9). We were also motivated by the issues raised by Shieh and Spector in [51], in particular their s Open Problem 1.10, which by (1.4) clearly holds when L˜A = (−∆) . Discussing this issue with examples, we give a counterexample in Example 2.9, showing how interesting is their Open Problem 1.10 for general strictly elliptic and bounded matrices A. Then, we conjecture in the Open Problem that the kernel kA corresponding to L˜A has the required property if and only if L˜A is approximately a constant multiple of the fractional Laplacian, up to small bounded perturbations. In Section 3, we make use of the comparison property to consider obstacle-type problems involving the bilinear form Ea, thereby considering the nonlocal obstacle problem for which we derive results similar to 1 the classical case in H0 (Ω) as in [55], [35] and [42], such as the weak maximum principle and comparison properties. Making use of convergence properties of the fractional derivatives when s % 1 to the classical derivatives, as already observed in [44], [8] and [15], we show that the solution of the fractional obstacle problem for L˜A converges to the solution of the classical case corresponding to s = 1. By considering the approximation of the obstacle problem by semilinear problems using a bounded penalization, in Section 4, we give a direct proof of the Lewy-Stampacchia inequalities for the obstacle- # type problems. Here we consider the non-homogeneous data F = f ∈ L2 (Ω) not only for the one obstacle problem but also for the two obstacles problem and for the N-membranes problem in the nonlocal framework, extending the results of [49]. In particular, we extend the estimates in energy of the difference between the approximating solutions and the solutions of the one and the two obstacle problems, which may be useful for numerical applications such as in [10] or [41]. More important is the use of the Lewy-Stampacchia inequalities that, upon restricting a to the symmetric case, allows the application of the results of [26] to obtain locally the H¨olderregularity of the solutions to those three nonlocal obstacle type problems. Such regularity results are weaker than those obtained from the fractional Laplacian (such as in [47]) or other commonly considered nonlocal kernels [27], since a is in general not a constant multiple of |x − y|−d−2s. In this special case, when s 2 d d La = (−∆) , the one obstacle problem can be written for f ∈ [L (R )]

s s s s u ∈ Kψ(Ω) : (D u − f) · D (v − u) dx ≥ 0 ∀v ∈ Kψ(Ω), ˆ d R as well as for the corresponding inequalities for the two obstacles and the N-membranes problems. We are then able to use the results of [9] together with the Lewy-Stampacchia inequalities to obtain locally 2s,p # 1 regular solutions in the fractional Sobolev space Wloc (Ω) for p ≥ 2 and also in C (Ω) for s > 1/2 and p > d/(2s − 1) when Ds · f ∈ Lp(Rd).

4 In Section 5, we further consider some properties related to the fractional s-capacity extending some classical results of Stampacchia [55]. We characterize the order dual of Hs(Ω) as the dual space of L2 (Ω), 0 Cs i.e. the space of quasi-continuous functions with respect to the s-capacity which are in absolute value quasi- s everywhere dominated by H0 (Ω) functions, extending results of [6]. That dual space corresponds then to the elements of H−s(Ω) that are also bounded measures, i.e. (L2 (Ω))0 = H−s(Ω) ∩ M(Ω). Therefore, using Cs the strict T-monotonicity of La, we state the Lewy-Stampacchia inequalities in this dual space. This section ends with some new remarks on the relations of the Ea obstacle problem and the s-capacity.

2 The Anisotropic Non-symmetric Nonlocal Bilinear Form 2.1 The nonlocal bilinear form as a coercive Dirichlet form s Our first main result is to show that the nonlocal bilinear form together with its domain, (Ea,H0 (Ω)), defined as

Ea(u, v) := P.V. v˜(x)(˜u(x) − u˜(y))a(x, y) dy dx, (2.1) ˆ d ˆ d R R s c d d whereu, ˜ v˜ are the zero extensions of u, v ∈ H0 (Ω) to Ω and a : R × R → [0, ∞), d ≥ 1 is a not necessarily symmetric kernel satisfying

2 d+2s ∗ 2 d a∗cd,s ≤ a(x, y)|x − y| ≤ a cd,s ∀x, y ∈ R , x 6= y, (2.2) is in fact a regular (not necessarily symmetric) Dirichlet form. This will also imply that the nonlocal bilinear form is also strictly T-monotone and will give us many properties, including Harnack’s inequality (see for example [11]), H¨olderregularity of solutions of equations involving this bilinear form (see for instance [32], [33], or [34]), and other results in stochastic processes (as given in [29]). We will begin with a remark on the symmetric case.

s d d Proposition 2.1. For given u, v ∈ H0 (Ω) and a : R × R → [0, ∞) symmetric, we have

v˜(x)(˜u(x) − u˜(y))a(x, y)χε(x, y) dy dx ˆ d ˆ d R R 1 = (˜u(x) − u˜(y))(˜v(x) − v˜(y))a(x, y)χε(x, y) dy dx (2.3) 2 ˆ d ˆ d R R

assuming that the integrands are summable for each fixed ε > 0, where we have set χε(x, y) as the charac- teristic function of the set {|x − y| > ε} for ε > 0.

∞ s Proof. We first show the result for u, v ∈ Cc (Ω), and extend it by density to u, v ∈ H0 (Ω). The integral term

J := v˜(x)(˜u(x) − u˜(y))a(x, y)χε(x, y) dy dx ˆ d ˆ d R R can also be written in the form, using the symmetry of a,

J = v˜(y)(˜u(y) − u˜(x))a(x, y)χε(x, y) dx dy ˆ d ˆ d R R Then, by Fubini theorem,

J = − v˜(y)(˜u(x) − u˜(y))a(x, y)χε(x, y) dy dx. ˆ d ˆ d R R Taking the sum of this and the first equation above, we obtain the result

2J = (˜v(x) − v˜(y))(˜u(x) − u˜(y))a(x, y)χε(x, y) dy dx. ˆ d ˆ d R R

5 Remark 2.2. The bilinear form (2.1) is also well-defined for u, v ∈ Hs(Rd). Indeed,

2 v(x)(u(x) − u(y)) lim a∗cd,s d+2s χε(x, y) dx dy ε→0 ˆ d ˆ d |x − y| R R

≤ lim v(x)(u(x) − u(y))a(x, y)χε(x, y) dx dy ε→0 ˆ d ˆ d R R ∗ 2 v(x)(u(x) − u(y)) ≤ lim a cd,s d+2s χε(x, y) dx dy ε→0 ˆ d ˆ d |x − y| R R with, applying Fubini’s theorem,

v(x)(u(x) − u(y)) lim d+2s χε(x, y) dy dx ε→0 ˆ d ˆ d |x − y| R R v(y)(u(y) − u(x)) = lim d+2s χε(x, y) dy dx ε→0 ˆ d ˆ d |x − y| R R 1 (v(x) − v(y)) (u(x) − u(y)) = d d dy dx 2 ˆ d ˆ d 2 +s 2 +s R R |x − y| |x − y| which is integrable by Cauchy-Schwarz inequality when u, v ∈ Hs(Rd). s It is well-known that H0 (Ω) is complete with respect to its norm, as in (1.5), therefore Ea is closed. s s s Furthermore, it can be shown that (Ea,H0 (Ω)) is regular, i.e. H0 (Ω) ∩ Cc(Ω) is dense in H0 (Ω) in the s H0 (Ω)-norm, and dense in Cc(Ω) with uniform norm. The first density result follows from the density s of compactly supported smooth functions in the space H0 (Ω). The second density result follows since ∞ s ∞ Cc (Ω) ⊂ H0 (Ω) and Cc (Ω) is dense in Cc(Ω) with uniform norm, by considering the mollification of any Cc(Ω) function. Recall that a coercive closed (not necessarily symmetric) bilinear form E on L2(Ω) is a Dirichlet form ([37] Proposition I.4.7 and equation (4.7) pages 34–35) if and only if the following property holds: For each ε > 0, there exists a real function φε(t), t ∈ R, such that

0 0 0 φε(t) = t, ∀t ∈ [0, 1] − ε ≤ φε(t) ≤ 1 + ε, ∀t ∈ R, 0 ≤ φε(t ) − φε(t) ≤ t − t whenever t < t (2.4) ( s s lim infε→0 E(φε(u), u − φε(u)) ≥ 0, u ∈ H0 (Ω) =⇒ φε(u) ∈ H0 (Ω), (2.5) lim infε→0 E(u − φε(u), φε(u)) ≥ 0.

A classic example of φε is the mollification of a cut-off function (see [29] Example 1.2.1). Specifically, consider a mollifier such as  − 1 γe 1−|x|2 for |x| < 1 j(x) = 0 otherwise, where γ is a positive constant such that j(x) dx = 1. ˆ|x|≤1 −1 −1 Set jδ(x) = δ j(δ x) for δ > 0. For any ε > 0, consider the function ψε(t) = ((−ε) ∨ t) ∧ (1 + ε) on R and set φε(t) = jδ ∗ ψε(t) for 0 < δ < ε. Then our choice of φε satisfies (2.4). Furthermore, it satisfies φε(t) = 1 + ε for t ∈ [1 + 2ε, ∞) and φε(t) = −ε for t ∈ (−∞, −2ε], and |φε(t)| ≤ |t| with tφε(t) ≥ 0. ∞ ∞ ∞ Moreover, φε(t) is infinitely differentiable, so for any u ∈ Cc (Ω), φε(u) ∈ Cc (Ω). Since Cc (Ω) is s s dense in H0 (Ω), we can extend by density any results obtained, thereby obtaining φε(u) ∈ H0 (Ω) for all s u ∈ H0 (Ω).

6 With this φε, recalling that a ≥ 0 (as in section II.2(d) of [37]), by Fatou’s lemma,

lim inf Ea(φε(u), u − φε(u)) ε→0

≥ P.V. lim inf[u(x) −^φε(u(x))][φε^(u(x)) − φε^(u(y))]a(x, y) dy dx ˆ d ˆ d ε→0 R R

≥ P.V. lim inf[u(x) −^φε(u(x))][−φε^(u(y))]a(x, y) dy dx (since φε(u)[u − φε(u)] ≥ 0) ˆ d ˆ d ε→0 R R n ≥ P.V. lim inf − [u(x) −^φε(u(x))](−ε)χ{u(x)<0} + 0χ{0≤u(x)≤1} ˆ d ˆ d ε→0 R R o + (−ε)(−1 − ε)χ{u(x)>1} a(x, y) dy dx = 0 since u − φε(u) ≤ 0 for u ≤ 0, φε(t) = t for t ∈ [0, 1], and u − φε(u) ≥ u − 1 − ε ≥ −ε for u ≥ 1 respectively. Similarly, taking lim infε→0 in   Ea(u − φε(u), φε(u)) = P.V. φε^(u(x)) [u(x) −^φε(u(x))] − [u(y) −^φε(u(y))] a(x, y) dx dy, ˆ d ˆ d R R we conclude that Ea is a Dirichlet form. Therefore, we obtained the following theorem:

s Theorem 2.3. (Ea,H0 (Ω)) is a closed, regular Dirichlet form, which is also bounded and coercive. s s Proof. It remains only to show that the bilinear form Ea : H0 (Ω) × H0 (Ω) → R is bounded

∗ 2 −d−2s Ea(u, v) ≤ a cd,s P.V. v˜(x)(˜u(x) − u˜(y))|x − y| dy dx ˆ d ˆ d R R 1 ∗ 2 −d−2s = a cd,s (˜u(x) − u˜(y))(˜v(x) − v˜(y))|x − y| dy dx 2 ˆ d ˆ d R R 1/2 1/2 2 ! 2 ! ∗ 2 1 (˜u(x) − u˜(y)) 1 (˜v(x) − v˜(y)) ≤ a cd,s d+2s dx dy d+2s dx dy 2 ˆ d ˆ d |x − y| 2 ˆ d ˆ d |x − y| R R R R ∗ s s = a kD ukL2(Ω)kD vkL2(Ω) ∗ s ≤ a kuk s kvk s ∀u, v ∈ H (Ω) H0 (Ω) H0 (Ω) 0 by the Cauchy-Schwarz inequality, and coercive

1 2 2 −d−2s s 2 2 Ea(u, u) ≥ a∗cd,s (˜u(x) − u˜(y)) |x − y| dy dx = a∗kD ukL2( d) = a∗kukHs(Ω) . 2 ˆ d ˆ d R 0 R R

From this theorem, we obtain that Ea possesses the property of unit contraction. Indeed, by Proposition 4.3 and Theorem 4.4 of [37], we have the following corollary.

Corollary 2.4. For the regular Dirichlet form Ea the following properties hold:

(a) the unit contraction acts on Ea, i.e. v := (0 ∨ u) ∧ 1 satisfies Ea(v, v) ≤ Ea(u, u);

(b) the normal contraction acts on Ea, i.e. suppose v satisfies

d |v(x) − v(y)| ≤ |u(x) − u(y)|, x, y ∈ R

d |v(x)| ≤ |u(x)|, x ∈ R , s then v ∈ H0 (Ω) and Ea(v, v) ≤ Ea(u, u).

7 This result in fact follows from the fact that Ea is a Dirichlet form, and holds even if it is not regular. Since Ea is a regular Dirichlet form, it is possible to consider truncations, so we can introduce the positive and negative parts of v v+ ≡ v ∨ 0 and v− ≡ −v ∨ 0 = −(v ∧ 0), and we have the Jordan decomposition of v given by

v = v+ − v− and |v| ≡ v ∨ (−v) = v+ + v−

and the useful identities u ∨ v = u + (v − u)+ = v + (u − v)+, u ∧ v = u − (u − v)+ = v − (v − u)+. s It is well-known that such operations are closed in H0 (Ω) for 0 < s ≤ 1.

2.2 The strict T-monotonicity of Ea Since the assumptions on the kernel a imply, in particular, that it is non-negative, we can easily prove the following important property.

s −s Theorem 2.5. Ea is also strictly T-monotone in the following sense: La : H0 (Ω) → H (Ω) defined by

hLau, vi = Ea(u, v), (2.6) satisfies + s + hLav, v i > 0 ∀v ∈ H0 (Ω) such that v 6= 0.

Proof. La is strictly T-monotone because

− + + − − Ea(v , v ) = P.V. v˜ (x)(˜v (x) − v˜ (y))a(x, y) dx dy ˆ d ˆ d R R = − P.V. v˜+(x)˜v−(y)a(x, y) dx dy ˆ d ˆ d R R ≤ 0,

since v+(x)v−(x) = 0 as v+ and v− cannot both be nonzero at the same point x, and v+, v−, a ≥ 0. d+2s 2 Therefore, since a(x, y)|x − y| ≥ a∗cd,s with a∗ > 0,

+ + hLAv, v i = Ea(v, v ) + + − + = Ea v , v − Ea v , v

s + 2 ≥ a∗ |D v | ˆ d R which is strictly greater than 0 if v+ 6= 0.

2.3 The fractional bilinear form as a nonlocal bilinear form s In this section, we consider the linear operator L˜A defined with the fractional derivatives D by the continuous fractional bilinear form ˜ s s s hLAu, vi := A(x)D u · D v dx, ∀u, v ∈ H0 (Ω) (1.8) ˆ d R for a matrix A with bounded and measurable coefficients. The integral in (1.8) is well defined and we are d d going to show that this bilinear form can be rewritten with a measurable kernel kA : R × R , d ≥ 1, as

EkA (u, v) := P.V. v˜(x)(˜u(x) − u˜(y))kA(x, y) dy dx, (2.7) ˆ d ˆ d R R ∞ c whereu, ˜ v˜ are the zero extensions of u, v ∈ Cc (Ω) to Ω .

8 Theorem 2.6. Given a matrix A : Rd → Rd×d with bounded and measurable coefficients, there exists a kernel kA(x, y) independent of u, v satisfying

s s A(x)D u(x) · D v(x) dx = P.V. v(x)(u(x) − u(y))kA(x, y) dy dx (2.8) ˆ d ˆ d ˆ d R R R ∞ d for all u, v ∈ Cc (R ), where kA(x, y) is given by

2 y − z z − x kA(x, y) = cd,s P.V. A(z) d+s+1 · d+s+1 dz for x 6= y. (2.9) ˆ d |y − z| |z − x| R

Proof. Expanding the fractional bilinear form, we have, setting for simplicity d as , R ´ ´ A(z)Dsu(z) · Dsv(z) dz ˆ  y − z   (x − z)  = c2 A(z) (u(y) − u(z)) dy · (v(x) − v(z)) dx dz d,s ˆ ˆ |y − z|d+s+1 ˆ |z − x|d+s+1   2 (y − z)χδ(y, z) (x − z)χε(x, z) = cd,s lim A(z)(u(y) − u(z)) · (v(x) − v(z)) χη(x, y) dx dy dz ˚ ε,δ,η→0 |y − z|d+s+1 |z − x|d+s+1

2 (y − z)χδ(y, z) (x − z)χε(x, z) = cd,s lim A(z)(u(y) − u(z)) · (v(x) − v(z)) χη(x, y) dx dy dz, ε,δ,η→0 ˚ |y − z|d+s+1 |z − x|d+s+1

where χη(x, y) is the characteristic function on the set {|x − y| > η} and similarly defined for χε and χδ. The limit can be exchanged with the integral by the Fubini and Lebesgue theorems because the integrand is Lebesgue integrable. Therefore,

A(z)Dsu(z) · Dsv(z) dz ˆ

2 (y − z)χδ(y, z) (x − z)χε(x, z) = cd,s lim A(z)(u(y) − u(z)) · v(x) χη(x, y) dx dy dz ε,δ,η→0 ˚ |y − z|d+s+1 |z − x|d+s+1

2 (y − z)χδ(y, z) (x − z)χε(x, z) − cd,s lim A(z)(u(y) − u(z)) · v(z) χη(x, y) dx dy dz ε,δ,η→0 ˚ |y − z|d+s+1 |z − x|d+s+1

2 (y − z)χδ(y, z) (x − z)χε(x, z) = cd,s lim A(z)(u(y) − u(z)) · v(x) χη(x, y) dx dy dz ε,δ,η→0 ˚ |y − z|d+s+1 |z − x|d+s+1   2 (y − z)χδ(y, z) (x − z)χε(x, z) − cd,s lim A(z)(u(y) − u(z)) · v(z) lim χη(x, y) dx dy dz ε,δ→0 ¨ |y − z|d+s+1 ˆ η→0 |z − x|d+s+1

2 (y − z)χδ(y, z) (x − z)χε(x, z) = cd,s lim A(z)(u(y) − u(z)) χη(x, y) · v(x) dx dy dz, (2.10) ε,δ,η→0 ˚ |y − z|d+s+1 |z − x|d+s+1

(x−z)χε(x,z) using the fact that |x−z|d+s+1 dx = 0 for all ε > 0. Note that we can exchange the limit in η with the integrals because the´ functions are Lebesgue integrable as |x − y| → 0.

9 ∞ d Adding and subtracting u(x), as u, v ∈ Cc (R ), we have

A(z)Dsu(z) · Dsv(z) dz ˆ

2 (y − z)χδ(y, z) (x − z)χε(x, z) = cd,s lim A(z)[(u(y) − u(x)) + (u(x) − u(z))] · v(x) χη(x, y) dx dy dz ε,δ,η→0 ˚ |y − z|d+s+1 |z − x|d+s+1

2 (y − z)χδ(y, z) (x − z)χε(x, z) = cd,s lim A(z)(u(y) − u(x)) · v(x) χη(x, y) dx dy dz ε,δ,η→0 ˚ |y − z|d+s+1 |z − x|d+s+1 "   # 2 (y − z)χδ(y, z) (x − z)χε(x, z) + cd,s lim A(z) · (u(x) − u(z))v(x) χη(x, y) dx dy dz ε,δ,η→0 ˆ ˆ |y − z|d+s+1 ˆ |z − x|d+s+1

2 (y − z)χδ(y, z) (x − z)χε(x, z) = cd,s lim A(z)(u(y) − u(x)) · v(x) χη(x, y) dx dy dz ε,δ,η→0 ˚ |y − z|d+s+1 |z − x|d+s+1 "   # 2 (y − z)χδ(y, z) (x − z)χε(x, z) + cd,s lim A(z) · lim(u(x) − u(z))v(x) χη(x, y) dx dy dz ε,δ→0 ˆ ˆ |y − z|d+s+1 ˆ η→0 |z − x|d+s+1

2 (y − z)χδ(y, z) (x − z)χε(x, z) = cd,s lim A(z)(u(y) − u(x)) · v(x) χη(x, y) dx dy dz, (2.11) ε,δ,η→0 ˚ |y − z|d+s+1 |z − x|d+s+1

(y−z)χδ (y,z) where we make use of the fact that |y−z|d+s+1 · f(z) dy = 0 for all δ > 0 for any finite function f(z). Once ´ (y−z)χδ (y,z) again, the limit in η can be interchanged with triple integrals, because the factor |y−z|d+s+1 is integrable for δ > 0. Also, the function

(x − z)χε(x, z) f(z) := lim (u(x) − u(z))χη(x, y)v(x) dx η→0 ˆ |z − x|d+s+1

is a finite function of z, because the integrand has a singularity only at x = z and we have introduced the characteristic function χε(x, z). Furthermore, the Lipschitz continuity of u guarantees that the singularity is removable, since we have the factor u(x) − u(z). This is also the reason why we used the first expression for the expansion of Dsu, rather than the second one, which will only give us a factor of u(x) when we add and subtract u(x). Therefore, we can take the limit η → 0, so that it is just a function of z. Next, we apply Fubini’s theorem, since the integrand is Lebesgue integrable for fixed ε, δ, η > 0. Therefore,

A(z)Dsu(z) · Dsv(z) dz ˆ

2 (y − z)χδ(y, z) (x − z)χε(x, z) = cd,s lim A(z)(u(y) − u(x)) · v(x) χη(x, y) dz dy dx. ε,δ,η→0 ˚ |y − z|d+s+1 |z − x|d+s+1

Finally, regarding this limit as a double limit, in η and separately in ε and δ, which exists, we can consider the iterated limit in the following form

A(z)Dsu(z) · Dsv(z) dz ˆ   2 (y − z)χδ(y, z) (z − x)χε(x, z) = lim (u(x) − u(y))v(x) cd,s lim A(z) · dz χη(x, y) dy dx η→0 ¨ ε,δ→0 ˆ |y − z|d+s+1 |z − x|d+s+1 where we may interpret the term in the parentheses as the Cauchy principal value about the singularities z = x and z = y, i.e. as a function in x, y defined for x 6= y, by y − z z − x k (x, y) = c2 P.V. A(z) · dz. (2.9) A d,s ˆ |y − z|d+s+1 |z − x|d+s+1

10 Remark 2.7. Note that in general, kA is neither translation nor rotation invariant, unlike the case for the −d−2s fractional Laplacian. In particular, kA may not have the form j(x − y)|x − y| . Therefore, the kernel kA may have relevance for non-homogeneous, non-isotropic and nonlocal problems. Even in the case for A being the constant coefficient matrix when kA is translation invariant, it may not be rotation invariant, unless if A is a constant multiple of the identity matrix. Remark 2.8. Suppose that the matrix A is given by αI for a strictly positive finite constant α and the identity matrix I. Then, by (1.4) and Proposition 2.1, L˜αI defined by (1.8) can also be defined by a symmetric bilinear form as in (2.8) with a kernel α given by αc2 α(x, y) = d,s |x − y|d+2s which is, up to a constant, the kernel of the fractional Laplacian and satisfies (2.2). 2 However, we observe that this representation may not be unique, and kαI may not be equal to αcd,s|x − y|−d−2s. Indeed, consider an unbounded nonzero L2-integrable function h(x) : Rd → R which integrates to 0 over Rd and has support outside Ω. Let a(x, y) be a kernel satisfying (2.2) and define a˜(x, y) = a(x, y) + h(x)h(y), which is possible since the kernel is defined over all (Rd × Rd)\{x = y} and the integrability of h means that ∞ La˜ is well defined. Since h = 0 by the construction of h and, for any u, v ∈ Cc (Ω), uh˜ = 0 since they have disjoint supports, we´ have ´

hLa˜u, vi = P.V. v˜(x)(˜u(x) − u˜(y))(a(x, y) + h(x)h(y)) dy dx ˆ d ˆ d R R = P.V. v˜(x)(˜u(x) − u˜(y))a(x, y) dy dx ˆ d ˆ d R R + P.V. u˜(x)˜v(x)h(x)h(y) dy dx − P.V. u˜(y)˜v(x)h(x)h(y) dy dx ˆ d ˆ d ˆ d ˆ d R R R R = P.V. v˜(x)(˜u(x) − u˜(y))a(x, y) dy dx ˆ d ˆ d R R     + P.V. u˜(x)˜v(x)h(x) h(y) dy dx − P.V. v˜(x)h(x) u˜(y)h(y) dy dx ˆ d ˆ d ˆ d ˆ d R R R R

= P.V. v˜(x)(˜u(x) − u˜(y))a(x, y) dy dx = hLau, vi ˆ d ˆ d R R This example gives a class of non-uniqueness for the representation of the kernel. There may be more similar classes, and it will be interesting to know a characterization for the equivalent class of kernels. Furthermore, even if the kernel a satisfies the compatibility condition (2.2), since h may change sign and may be unbounded, our construction of the kernel a˜ may not satisfy the compatibility condition (2.2), nor the weaker compatibility conditions

( d+2s ∗ a∗ ≤ a(x, y)|x − y| ≤ a , |x − y| ≤ 1 0 (2.12) a(x, y) ≤ M|x − y|−d−s , |x − y| > 1

∗ 0 for some a∗, a , M, s > 0, as given in equation (1.11) of [51].

However, (2.2) is not satisfied for the kernel kA for a general matrix A. Indeed, we can construct a numerical counterexample as follows. Example 2.9. In d = 1, suppose s = 0.8, and consider the matrix A = α(x) where α(x) = 0.01 + 50H(x) for the smooth approximation H of the characteristic function of the interval [1, 1.5], such that H(x) = 1 in [1, 1.5] and less than 0.0001 outside [0.9, 1.6]. Then 0.5 − z z + 0.5 0.5 − z z + 0.5 k (−0.5, 0.5) = 0.01c2 P.V. dz + 50c2 P.V. H(z) dz. A 1,0.8 ˆ |0.5 − z|2.8 |z + 0.5|2.8 1,0.8 ˆ |0.5 − z|2.8 |z + 0.5|2.8 R R

11 Observe that the function 0.5 − z z + 0.5 κ (−0.5, 0.5, z) := 1,0.8 |0.5 − z|2.8 |z + 0.5|2.8 has the shape

Created in Symbolab is integrable but not absolutely integrable, and is strictly increasing and strictly negative in the interval [0.9, 1.6] with values (computed in Wolfram Alpha)

κ (−0.5, 0.5, 0.9) = −2.839, κ (−0.5, 0.5, 1.5) = −0.287, κ (−0.5, 0.5, z) dz ≈ 30 < 100. 1,0.8 1,0.8 ˆ 1,0.8 R Then, computing (2.9) for 0.5 − z z + 0.5 0.5 − z z + 0.5 k (−0.5, 0.5) = 0.01c2 P.V. dz + 50c2 P.V. H(z) dz A 1,0.8 ˆ |0.5 − z|2.8 |z + 0.5|2.8 1,0.8 ˆ |0.5 − z|2.8 |z + 0.5|2.8 R R 1.5 2 2 0.5 − z z + 0.5 < 0.01c1,0.8(100) + 50c1,0.8P.V. 2.8 2.8 dz ˆ1 |0.5 − z| |z + 0.5| 0.5 − z z + 0.5 + 50(0.0001)c2 P.V. dz 1,0.8 ˆ |0.5 − z|2.8 |z + 0.5|2.8 R\[0.9,1.6] 1.5 2 2 0.5 − z z + 0.5 = c1,0.8 + 50c1,0.8P.V. 2.8 2.8 dz ˆ1 |0.5 − z| |z + 0.5| ! 0.5 − z z + 0.5 1.6 0.5 − z z + 0.5 + 0.005c2 P.V. dz − P.V. dz 1,0.8 ˆ |0.5 − z|2.8 |z + 0.5|2.8 ˆ |0.5 − z|2.8 |z + 0.5|2.8 R 0.9 1.5 2 2 < c1,0.8 + 50c1,0.8P.V. κ1,0.8(−0.5, 0.5, 1.5) dz ˆ1 1.6 ! + 0.005c2 P.V. κ (−0.5, 0.5, z) dz − P.V. κ (−0.5, 0.5, 0.9) dz 1,0.8 ˆ 1,0.8 ˆ 1,0.8 R 0.9 2 2 2 2 < c1,0.8 + 50c1,0.8(−0.28)(0.5) + 0.005c1,0.8(100 − (−2.84)(1.6 − 0.9)) = −5.49c1,0.8 < 0 which contradicts (2.2). Compare with Open Problem 1.10 of [51].

The Theorem 2.6 and the last two remarks were inspired by the Open Problem 1.10 of [51], which asked if, given a symmetric matrix A satisfying (2.13), it is possible to find a kernel kA satisfying (2.12) such that (2.8) holds. Complementing this open problem, we propose the following conjecture (see also Remark 3.1 of [20]):

Open Problem. Suppose A : Rd → Rd×d is a bounded, measurable and strictly elliptic matrix such that 2 ∗ ∗ ∗ a∗|z| ≤ A(x)z · z and A(x)z · z ≤ a |z||z | (2.13)

∗ d ∗ d for some a∗, a > 0 for all x ∈ R and all z, z ∈ R . Let kA be a corresponding kernel which is continuous outside the diagonal x = y and satisfies

s s A(x)D u(x) · D v(x) dx = P.V. v(x)(u(x) − u(y))kA(x, y) dy dx ˆ d ˆ d ˆ d R R R

12 ∞ d for all u, v ∈ Cc (R ). Then, there exists an equivalent kernel kA satisfying

d+2s ∗ d a∗ ≤ kA(x, y)|x − y| ≤ a ∀x, y ∈ R , x 6= y

∗ for some a∗, a > 0 if and only if A is a bounded small perturbation α˜(x) of the identity matrix (up to a I positive constant α), i.e. A = (α +α ˜(x)) for some strictly positive finite constant α >> supx |α˜(x)|.

3 The Nonlocal and the Fractional Obstacle Problem

We start by considering the for F ∈ H−s(Ω) defined by

s hF, vi = f#v + f · D v ˆ ˆ d Ω R

s 2 d d 2# # 2d d for v ∈ H0 (Ω), f = (f1, . . . , fd) ∈ [L (R )] and f# ∈ L (Ω) where 2 = d+2s when s < 2 , and if d = 1, # 1 # 1 2 = q for any finite q > 1 when s = 2 and 2 = 1 when s > 2 . By the Riesz representation theorem, we have s s s s s s ∃!φ ∈ H0 (Ω) : D φ · D v = h−D · D φ, vi = hF, vi, ∀v ∈ H0 (Ω). ˆ d R −s s s s 2 d d Therefore, F ∈ H (Ω) may be given by F = −D · D φ = −D · g for some g = (g1, . . . , gd) ∈ [L (R )] and, by the Sobolev-Poincar´einequality, it satisfies

kgk 2 d d =kF k −s ≤ CS f# # +kfk 2 d d . [L (R )] H (Ω) L2 (Ω) [L (R )] In order for s F ∼ f# − D · f −s to lie in the positive cone of H (Ω), it is enough for f# to be non-negative almost everywhere in Ω and Ds · f ≤ 0 in the distributional sense in Rd. We next recall the following useful inequalities.

d Lemma 3.1 (Sobolev-Poincar´einequality, Theorem 1.8 of [50]). Let s ∈ (σ, 1) such that s < 2 and σ > 0. Then there exists a constant CS = C(d, s) > 0 such that

s kuk 2∗ ≤ CSkD uk 2 d L (Ω) L (R )

s ∗ 2d for all u ∈ H0 (Ω), where 2 = d−2s > 0. Lemma 3.2 (Fractional Poincar´einequality, Theorem 2.9 of [8]). Let s ∈ (0, 1). Then there exists a constant CP = C(d, Ω)/s > 0 such that s kuk 2 ≤ CP kD uk 2 d L (Ω) L (R ) s for all u ∈ H0 (Ω).

3.1 The nonlocal obstacle problem and its properties s As a consequence of the properties of the bilinear form Ea defined by (2.1) in H0 (Ω) for 0 < s < 1, we can derive classical properties of the fractional obstacle problems, following most of the approach of Section 4:5 in [42].

d Theorem 3.3 (Obstacle Problem). Let Ω ⊂ R be a open bounded domain with Lipschitz boundary, f# ∈ # L2 (Ω), f ∈ [L2(Rd)]d, and a : Rd × Rd → [0, ∞) be strictly elliptic, bounded and measurable as in (2.2). Then, for every function ψ, measurable in Ω, and admissible in the sense that the closed

s s Kψ = {v ∈ H0 (Ω) : v ≥ ψ a.e. in Ω}= 6 ∅,

13 s there exists a unique u ∈ Kψ such that

s s Ea(u, v − u) ≥ f#(v − u) + f · D (v − u) ∀v ∈ Kψ. (3.1) ˆ ˆ d Ω R Moreover, suppose F, Fˆ are given as in the beginning of this section for two different obstacle problems defined in (3.1), then the solution map F 7→ u is Lipschitz continuous, i.e.   1 ˆ ˆ ku − uˆk s ≤ CS f# − f# + f − f . H0 (Ω) 2# 2 d d a∗ L (Ω) [L (R )]

s Proof. This is just a direct application of the Stampacchia theorem, since the bilinear form Ea : H0 (Ω) × s H0 (Ω) → R is bounded and coercive by Theorem 2.3. For the continuous dependence on data, if u, uˆ are the solutions corresponding to different data F and Fˆ for the obstacle problem respectively, we set v =u ˆ in the inequality for u and v = u in the inequality for uˆ, and take the difference to obtain

ˆ ˆ s Ea(u − u,ˆ u − uˆ) ≤ (f# − f#)(u − uˆ) + (f − f) · D (u − uˆ). ˆ ˆ d Ω R By the fractional Sobolev inequality and the Cauchy-Schwarz inequality, we have

s 2 a∗ D (u − uˆ) 2 d ≤ Ea(u − u,ˆ u − uˆ) L (R ) ˆ ˆ s ≤ (f# − f#)(u − uˆ) + (f − f) · D (u − uˆ) ˆ ˆ d Ω R s ˆ ∗ ˆ ≤ f# − f# ku − uˆk 2 + f − f D (u − uˆ) 2 d d 2# L (Ω) 2 d d [L ( )] L (Ω) [L (R )] R   ˆ ˆ s ≤ CS f# − f# + f − f D (u − uˆ) 2 d d 2# 2 d d [L ( )] L (Ω) [L (R )] R where CS is the Sobolev constant of Lemma 3.1. Furthermore, we have the following properties of the solution as in the classical case, making use of the strict T-monotonicity of Ea and the fractional Poincar´einequality. See for example, Chapter IV of [35] or Section 4:5–6 of [42], where the proofs can be transferred to the nonlocal case almost in the same manner. Proposition 3.4. (i) (Comparison principle). Suppose u is the solution of the (3.1) with data F and convex set s , and uˆ be the solution with data Fˆ and convex set s . If ψ ≥ ψˆ and Kψ Kψˆ F ≥ Fˆ, then u ≥ uˆ a.e. in Ω. (ii) (Weak maximum principle). In the obstacle problem (3.1), one has

u ≥ 0 a.e. in Ω, if F ≥ 0; and

u ≤ 0 ∨ sup ψ a.e. in Ω, if F ≤ 0. Ω

s (iii) (Complementary problem). When ψ ∈ H0 (Ω), the variational problem (3.1) is equivalent to the non- linear complementary problem

u ≥ ψ, Lau − F ≥ 0 and hLau − F, u − ψi = 0. (3.2)

(iv) (Comparison of supersolutions). (a) If u is the solution to the variational inequality (3.1) and w is any supersolution, then u ≤ w; (b) If v and w are two supersolutions to the variational inequality (3.1), then v ∧ w is also a super- solution,

14 s where a supersolution is an element w ∈ H0 (Ω) satisfying w ≥ ψ and Law − F ≥ 0 in the sense of order dual.

s (v) The solution u of the obstacle problem (3.1) is the unique function in H0 (Ω), such that,

s u = min{w ∈ H0 (Ω) : Law − F ≥ 0, w ≥ ψ in Ω}.

(vi)( L∞ estimates) The following estimate holds: ˆ ku − uˆkL∞(Ω) ≤ kψ − ψkL∞(Ω). Similarly, as in Theorem 6.1 of Chapter 4 of [42], we can prove the following additional result for the Dirichlet form Eλ with λ > 0.

Proposition 3.5. All the results in Proposition 3.4 holds for Eλ when λ > 0, where

s Eλ(u, v) = Ea(u, v) + λ uv, u, v ∈ H0 (Ω). (3.3) ˆΩ Moreover in this case, when f ≡ 0, the following maximum principle holds a.e. in Ω: f  f  0 ∧ inf # ≤ u ≤ 0 ∨ sup ψ ∨ sup # . Ω λ Ω Ω λ

d d Remark 3.6. If Ω = R and since the kernel a is defined in the whole R , the domain of Eλ, D(Eλ), is s d instead given by H (R ), and the Dirichlet form Eλ is coercive for λ > 0.

3.2 Convergence of the fractional s-obstacle problem as s % 1 We start with a continuous dependence property of the Riesz derivatives as s varies.

s0 s 2 d d 0 0 Lemma 3.7. For u ∈ H0 (Ω), D u is continuous in [L (R )] as s varies in [σ, s ] for 0 < σ < s ≤ 1. As a consequence, we have the following estimate: for σ ≤ s ≤ 1,

σ s kD uk 2 d ≤ cσkD uk 2 d , (3.4) L (R ) L (R ) s for any u ∈ H0 (Ω), where the constant cσ is independent of s. ∞ s Proof. Consider first u ∈ Cc (Ω). Recall that the Fourier transform of D u is given, by Theorem 1.4 of [50],

Ddsu = (2π)siξ|ξ|−1+su,ˆ

s0 whereu ˆ is the Fourier transform of u extended by 0 outside Ω. Since u ∈ H0 (Ω), the mapping s 7→ −(2π)siξ|ξ|−1+suˆ is continuous in [L2(Rd)]d as s varies in [σ, s0]. Therefore, conducting the inverse Fourier s t 0 ∞ transform, we have lims→t D u − D u 2 d = 0 for t ∈ [σ, s ] for u ∈ Cc (Ω). Extending this by density to L (R ) s0 all u ∈ H0 (Ω), we have the continuity result on s. Finally, the estimate (3.4) follows similarly to Proposition 2.7 of [8] using Lemma 3.2. As it could be expected, we have a continuous transition from the fractional obstacle problem to the classical local obstacle problem as s % 1 in the following sense. We now consider the obstacle problem

1 1 s s Theorem 3.8. Suppose ψ is such that Kψ := {v ∈ H0 (Ω) : v ≥ ψ a.e. in Ω} 6= ∅. Let u ∈ Kψ for 0 < s < 1 be the solution to the fractional obstacle problem, i.e.

s s s s s s AD u · D (v − u ) ≥ hF, v − u i ∀v ∈ Kψ, ˆ d R where A : Rd → Rd×d is a bounded, measurable and strictly elliptic matrix satisfying 2 ∗ ∗ ∗ a∗|z| ≤ A(x)z · z and A(x)z · z ≤ a |z||z |, (2.13)

15 −σ s s s and F ∈ H (Ω). Then, there exists a unique solution u ∈ H0 (Ω). Furthermore, the sequence (u )s σ 1 converges strongly to u in H0 (Ω) as s % 1 for any fixed 0 < σ < 1, where u ∈ Kψ solves uniquely the obstacle problem for s = 1, i.e.

1 ADu · D(v − u) ≥ hF, v − ui ∀v ∈ Kψ. ˆΩ Proof. The existence of a solution follows from a direct application of the Stampacchia theorem, since the bilinear form s s hL˜Au, vi = AD u · D v ˆ d R is bounded ˜ ∗ s hLAu, vi ≤ a kuk s kvk s ∀u, v ∈ H (Ω) H0 (Ω) H0 (Ω) 0 and coercive ˜ s 2 2 hLAu, vi ≥ a∗kD uk 2 d = a∗kuk s . L (R ) H0 (Ω) For uniqueness, if u, uˆ are the solutions corresponding to the same data F for the obstacle problem, we set v =u ˆ in the inequality for u and v = u in the inequality foru ˆ, and take the difference to obtain

s 2 s s s a∗ D (u − uˆ) L2( d) ≤ AD (u − uˆ) · D (u − uˆ) ≤ (f# − f#)(u − uˆ) + (f − f) · D (u − uˆ) = 0, R ˆ d ˆ ˆ d R Ω R so u ≡ uˆ. Next, we want to show the convergence of the fractional obstacle problem to the classical one. We first s s 1 1 T s prove an a priori estimate for D u . For v ∈ Kψ = σ≤s<1 Kψ, by Cauchy-Schwarz inequality and Sobolev inequality,

s s 2 s s s s a∗kD u kL2( d) ≤ AD u · D u R ˆ d R ≤ ADsus · Dsv1 − hF, v1 − usi ˆ d R ∗2 2 0 a  s s 2 1 s 1 1 1 2  s s 2 ≤ kD u k 2 d + D v − hF, v i + kF k −s + kD u k 2 d L (R ) 2 d 0 H (Ω) L (R ) 2 2 L (R ) 2 2 ∗2 2 2 a∗ s s 2 a s 1 1 cσ 2 a∗ s s 2 ≤ kD u k 2 d + D v − hF, v i + kF k −σ + kD u k 2 d L (R ) 2 d H (Ω) L (R ) 4 a∗ L (R ) a∗ 4

a∗ 0 a∗ 2 by taking  = 2a∗2 and  = 2 and cσ may be chosen independent of s for 0 < s ≤ 1, as a consequence of (3.4) for the dual norms k·kH−s(Ω). Therefore, we have

s s kD u k 2 d ≤ C, L (R ) ∗ where the constant C = C(σ, a∗, a ) > 0 is independent of s ≥ σ. Also by (3.4), for σ ≤ s < 1, we have

σ s s s kD u k 2 d ≤ cσkD u k 2 d ≤ C, (3.5) L (R ) L (R ) for some constant C independent of s, σ ≤ s < 1, and we may take a sequence

s s 2 d d D u −−−* ζ in [L (R )] -weak s%1

s σ for some ζ. By compactness, since u is also uniformly bounded in H0 (Ω), there exists a subsequence and a limit u ∈ L2(Ω) such that us −−−→ u strongly in L2(Ω). s%1 ∞ d ˜ Now, by Lemma 3.7, for all Φ ∈ [Cc (Ω)] , denoting by Φ the zero extension of Φ outside Ω, s 2 d d D · Φ˜ → D · Φ˜ in [L (R )] ,

16 therefore Dsus · Φ˜ = − u˜s(Ds · Φ)˜ −−−→ − u˜(D · Φ)˜ . ˆ d ˆ d s%1 ˆ d R R R But by the a priori estimate on Dsus,

s s ˜ D u · Φ ≤ CkΦk[L2(Ω)]d , ˆ d R which implies that

˜ u˜(D · Φ) ≤ CkΦk[L2(Ω)]d . ˆ d R This means that Du˜ ∈ [L2(Rd)]d, and since Ω has a Lipschitz boundary, Duf = Du˜, so Du ∈ [L2(Ω)]d. σ 1 Together with the first inequality in (3.4) which implies that u ∈ Kψ for any σ < 1, we have u ∈ Kψ. Furthermore, by Lemma 3.7, Dsu → Duf strongly in [L2(Rd)]d as s % 1, so

Ds(us − u) · Φ˜ = − (u^s − u)(Ds · Φ) → 0, ˆ d ˆ d R R therefore ζ = lim Dsus = Du.f s%1 1 s Finally, it remains to show that u satisfies the obstacle problem for s = 1. For any v ∈ Kψ ⊂ Kψ, since Dsus are uniformly bounded, we have, up to a subsequence and using the lower-semicontinuity of 1 T Asym = 2 (A + A ),

ADu · Dv = ADuf · Dvf ˆ ˆ d Ω R = lim ADsus · Dsv s%1 ˆ d R ≥ limhF, v − usi + lim inf ADsus · Dsus s%1 s%1 ˆ d R s s s s = hF, v − ui + lim inf AsymD u · D u s%1 ˆ d R

≥ hF, v − ui + AsymDuf · Duf ˆ d R = hF, v − ui + ADu · Du ˆΩ since Dsus * Duf weakly in [L2(Rd)]d and Dsv → Dvf strongly in [L2(Rd)]d. The conclusion follows by the σ σ0 0 compactness of the inclusion of H0 (Ω) in H0 (Ω) when σ > σ .

Remark 3.9. The case with A = I corresponds to the obstacle problem for the fractional Laplacian and was first considered by Silvestre in [54]. Indeed, from (1.4), since 2 s s cd,s (u(x) − u(y))(v(x) − v(y)) D u · D v = d+2s dx dy, ˆ d 2 ˆ d ˆ d |x − y| R R R s s also holds for u, v ∈ H0 (Ω), Theorem 3.8 gives the convergence of the solution u to the nonlocal obstacle problem (3.1), which is equivalent, up to a constant, to

s s s s s s s s u ∈ Kψ : D u · D (v − u ) ≥ hF, v − u i ∀v ∈ Kψ ˆ d R towards the solution u of the classical problem

1 1 u ∈ Kψ : Du · D(v − u) ≥ hF, v − ui ∀v ∈ Kψ. ˆΩ

Remark 3.10. A similar convergence result as s % 1 remains open for the general nonlocal operator La.

17 4 Lewy-Stampacchia Inequalities and Local Regularity

In this section, we take f = f# and f = 0. We give a direct proof of the Lewy-Stampacchia inequalities. This will follow much of the approach of Section 5:3.3 in [42] or Chapter IV of [35]. The Lewy-Stampacchia inequalities will allow us to apply the results of [26][28] to obtain local H¨olderregularity of the solutions s when a is symmetric, and additional regularity on fractional Sobolev spaces when La = (−∆) using [9].

s 4.1 Bounded penalization of the obstacle problem in H0 (Ω) s d −s Assume now that the obstacle ψ ∈ H (R ), so that we may define Laψ ∈ H (Ω) by (1.7) for any test s s function v ∈ H0 (Ω), and ψ is such that the convex set Kψ 6= ∅. Consider the approximation to the obstacle problem, where the penalization is based on any nondecreasing Lipschitz function θ : R → [0, 1] such that

0,1 0 θ ∈ C (R), θ ≥ 0, θ(+∞) = 1 and θ(t) = 0 for t ≤ 0;

∃Cθ > 0 : [1 − θ(t)]t ≤ Cθ, t > 0. t  Then, for any ε > 0, consider the family of functions θε(t) = θ ε , t ∈ R, which converges as ε → 0 to the multi-valued Heaviside graph. Examples of such sequences of functions include θ(t) = t/(1 + t), θ(t) = (2/π) arctan t, or from any non-decreasing Lipschitz function 0 ≤ θ ≤ 1 such that θ(t) = 1 for t ≥ t∗ > 0. Assume that + 2# f, (Laψ − f) ∈ L (Ω). (4.1)

# For ζ ∈ L2 (Ω) such that + ζ ≥ (Laψ − f) a.e. in Ω, consider now the one parameter family of approximating semilinear problems in variational form

s s uε ∈ H0 (Ω) : Ea(uε, v) + ζθε(uε − ψ)v = (f + ζ)v ∀v ∈ H0 (Ω). (3.1ε) ˆΩ ˆΩ Arguing as in the proof of Theorem 5:3.1 of [42], we have the following theorem.

s Theorem 4.1. The unique solution uε of the semilinear boundary value problem (3.1ε) is such that uε ∈ Kψ for each ε > 0, and it defines a monotone decreasing sequence, converging as ε → 0 to the solution u of the obstacle problem (3.1) with the error estimate

2 ku − uεk s ≤ ε(Cθ/a∗)kζk 1 . (4.2) H0 (Ω) L (Ω)

Remark 4.2. (i) We can formally interpret the variational equation (3.1ε) as corresponding to the fol- lowing semilinear boundary value problem:

d Lauε + ζθε(uε − ψ) = f + ζ in Ω, uε = 0 on R \Ω.

(ii) One can also consider the translated penalization, given by

 t   t  θ¯ (t) = 1 − θ − = θ¯ − ε ε ε

for t ∈ R and ε > 0, to approach the solution of the obstacle problem from below using monotonicity. Then we have that the unique solution u¯ε of the penalized problem

s ¯ s ∼ u¯ε ∈ H0 (Ω) : Ea(¯uε, v) + ζθε(¯uε − ψ)v = (f + ζ)v ∀v ∈ H0 (Ω). (3.1ε ) ˆΩ ˆΩ defines a monotone increasing sequence converging to the solution u of the obstacle problem (3.1) weakly s in H0 (Ω).

18 (iii) For special choices of the function θ, we can estimate the uniform convergence for the approximations of u by penalization. Suppose that, in addition to the conditions on θ,

θ(t) = 1 for t ≥ 1,

then the approximating solution uε of (3.1ε) verifies, for each ε > 0,

uε − ε ≤ u ≤ uε a.e. in Ω.

If, additionally, θ verifies t ≤ θ(t) ≤ 1 if 0 ≤ t ≤ 1,

the approximating solution u¯ε yields

u¯ε ≤ u ≤ uε and 0 ≤ uε − u¯ε ≤ ε a.e. in Ω.

From this theorem, we can derive the Lewy-Stampacchia inequality. Theorem 4.3 (Lewy-Stampacchia inequality). Under the assumptions (4.1), the solution u of the obstacle problem satisfies f ≤ Lau ≤ f ∨ Laψ a.e. in Ω. 2# In particular, Lau ∈ L (Ω).

+ Proof. Choosing ζ = (Laψ − f) in (3.1ε), and making use of the property of θ that 0 ≤ 1 − θε ≤ 1, then s for any ε > 0 and any v ∈ H0 (Ω), v ≥ 0, we have

+ + Lauεv = [f + (Laψ − f) (1 − θε)]v ≤ [f + (Laψ − f) ]v ˆΩ ˆΩ ˆΩ from the variational form, and on the other hand

Lauε = f + ζ − ζθε = f + ζ(1 − θε) ≥ f holds a.e. in Ω. Together, these give

+ fv ≤ Lauεv ≤ [f + (Laψ − f) ]v = [f ∨ (Laψ)]v. ˆΩ ˆΩ ˆΩ ˆΩ Letting ε → 0, this holds for u. Since v is arbitrary, we have the result. Remark 4.4. The Lewy-Stampacchia inequalities for nonlocal obstacle problems have been first obtained in [49] for a class of symmetric integrodifferential operators LK , with even kernels K, which are also strictly ∞ T-monotone and include the fractional Laplacian, and with f and LK ψ ∈ L (Ω).

4.2 Multiple obstacles problem 4.2.1 Two obstacles problem We next consider the two obstacles problem with a Dirichlet boundary condition in a bounded domain d s Ω ⊂ R with Lipschitz boundary, which consists of finding u ∈ Kψ,ϕ such that

s Ea(u, v − u) ≥ f(v − u), ∀v ∈ Kψ,ϕ, (4.3) ˆΩ

# where f ∈ L2 (Ω) and s s Kψ,ϕ = {v ∈ H0 (Ω) : ψ ≤ v ≤ ϕ a.e. in Ω}. (4.4) s s d Assume that ϕ and ψ are measurable and admissible obstacles in Ω such that Kψ,ϕ 6= ∅. When ϕ, ψ ∈ H (R ), a sufficient condition for these two assumptions to hold is to assume ϕ ≥ ψ a.e. in Ω and ϕ ≥ 0 ≥ ψ a.e. in Ωc.

19 Theorem 4.5. The two obstacles problem has a unique solution. Moreover, if u = u(f, ϕ, ψ) and uˆ = u(f,ˆ ϕ,ˆ ψˆ) are solutions in s and in s , respectively, of the two obstacles problem, then Kψ,ϕ Kψ,ˆ ϕˆ

f ≥ f,ˆ ϕ ≥ ϕ,ˆ ψ ≥ ψˆ implies u ≥ uˆ a.e. in Ω.

In addition, if f = fˆ, we have the L∞ estimate ˆ ku − uˆkL∞(Ω) ≤ kψ − ψkL∞(Ω) +kϕ − ϕˆkL∞(Ω) . Proof. The existence and uniqueness follows as, in the previous sections, from the monotonicity, coercivity, continuity and boundedness of the operator La and the Stampacchia theorem. The comparison property follows also as previous by the T-monotonicity of La. The L∞ estimate follows as well, as in the classical one obstacle problem.

Corresponding to the two obstacles problem, we also have the Lewy-Stampacchia inequality.

2# Theorem 4.6. The solution u of the two obstacles problem, for f, Laϕ, Laψ ∈ L (Ω) such that ϕ, ψ ∈ s d H (R ) are compatible and Laϕ, Laψ are given by (1.7), satisfies

f ∧ Laϕ ≤ Lau ≤ f ∨ Laψ a.e. in Ω, (4.5)

2# and therefore Lau ∈ L (Ω). Proof. The proof is similar to that of the classical case s = 1, now for two obstacles. Consider the penalized problem given by

s s uε ∈ H0 (Ω) : hLauε, vi + ζψθε(uε − ψ)v − ζϕθε(ϕ − uε)v = (f + ζψ − ζϕ)v ∀v ∈ H0 (Ω) ˆΩ ˆΩ ˆΩ for + − ζψ ≥ (Laψ − f) , ζϕ ≥ (Laϕ − f) , s with θε(t) = 1 for t ≥ ε. Then, there is a unique solution uε ∈ H0 (Ω) such that ψ ≤ uε ≤ ϕ + ε for each ε > 0. Indeed, we obtain the existence and uniqueness of the solution by the Stampacchia theorem as before. To show that ψ ≤ uε ≤ ϕ + ε, we have

+ s hLaψ, vi ≤ h(Laψ − f) + f, vi ≤ (ζψ + f)v ∀v ∈ H0 (Ω), v ≥ 0, and ˆΩ

− s hLaϕ, vi = h(Laϕ − f) + f, vi ≥ hf − (Laϕ − f) , vi ≥ (f − ζϕ)v ∀v ∈ H0 (Ω), v ≥ 0. ˆΩ + s Taking v = (ψ − uε) ∈ H0 (Ω) and using the strict T-monotonicity of La, we have

+ 2 + + a∗ (ψ − uε) s ≤Ea(ψ, (ψ − uε) ) − Ea(uε, (ψ − uε) ) H0 (Ω) + + ≤ (f + ζψ)(ψ − uε) − {f + ζψ[1 − θε(uε − ψ)] − ζϕ[1 − θε(ϕ − uε)]}(ψ − uε) ˆΩ ˆΩ + + = ζψθε(uε − ψ)(ψ − uε) − {ζϕ[1 − θε(ϕ − uε)]}(ψ − uε) ˆΩ ˆΩ ≤0

because the first term is non-positive, while the factors in the second term are all non-negative. Therefore,

20 + s uε ≥ ψ. Similarly, taking v = (uε − ϕ − ε) ∈ H0 (Ω) gives

+ 2 s + 2 a∗ (uε − ϕ − ε) s = a∗ D (uε − ϕ − ε) 2 d H0 (Ω) L (R ) + + ≤ Ea((uε − ϕ − ε) , (uε − ϕ − ε) ) + ≤ Ea(uε − ϕ − ε, (uε − ϕ − ε) ) + ≤ Ea(uε − ϕ, (uε − ϕ − ε) )  + +  = Ea(uε, (uε − ϕ − ε) ) − Ea(ϕ, (uε − ϕ − ε) )

+ ≤ {f + ζψ[1 − θε(uε − ψ)] − ζϕ[1 − θε(ϕ − uε)]}(uε − ϕ − ε) ˆΩ + − (f − ζϕ)(uε − ϕ − ε) ˆΩ + + = {ζψ[1 − θε(uε − ψ)]}(uε − ϕ − ε) + [ζϕθε(ϕ − uε)](uε − ϕ − ε) ˆΩ ˆΩ + = ζψ[1 − θε(uε − ψ)](uε − ϕ − ε) ˆΩ + ≤ ζψ[1 − θε(uε − ψ)](uε − ψ − ε) since ϕ ≥ ψ ˆΩ = 0

Therefore, uε ≤ ϕ + ε. Now, we can show that uε → u, so ψ ≤ uε ≤ ϕ + ε converges to ψ ≤ u ≤ ϕ. Take v = w − uε in the s penalized problem above for arbitrary w ∈ Kψ,ϕ, then

Ea(uε, w − uε) = f(w − uε) + ζψ[1 − θε(uε − ψ)](w − uε) − ζϕ[1 − θε(ϕ − uε)](w − uε) ˆΩ ˆΩ ˆΩ

≥ f(w − uε) + ζψ[1 − θε(uε − ψ)](ψ − uε) − ζϕ[1 − θε(ϕ − uε)](ϕ − uε) ˆΩ ˆΩ ˆΩ uε − ψ ϕ − uε = f(w − uε) − ε ζψ[1 − θε(uε − ψ)] − ε ζϕ[1 − θε(ϕ − uε)] ˆΩ ˆΩ ε ˆΩ ε

≥ f(w − uε) − εCθ (ζψ + ζϕ). ˆΩ ˆΩ

s Now, taking w = u ∈ Kψ,ϕ, we obtain

Ea(uε, u − uε) ≥ f(u − uε) − εCθ (ζψ + ζϕ), ˆΩ ˆΩ

s but taking v = uε ∈ Kψ in the original obstacle problem (3.1), we have

Ea(u, uε − u) ≥ f(uε − u). ˆΩ

Taking the difference of these two equations, by the linearity of Ea, we have

Ea(uε − u, uε − u) ≤ εCθ (ζψ + ζϕ). ˆΩ Using the ellipticity of a, we have

s 2 2 εCθ (ζψ + ζϕ) ≥ Ea(uε − u, uε − u) ≥ a∗ D (uε − u) L2( d) = a∗kuε − ukHs(Ω) . ˆΩ R 0

s Therefore, uε → u in H0 (Ω).

21 + − Choosing ζψ = (Laψ −f) and ζϕ = (Laϕ−f) , and making use of the property of θ that 0 ≤ 1−θε ≤ 1, s then for any ε > 0 and any v ∈ H0 (Ω), v ≥ 0, we have

+ − + (Lauε)v = [f + (Laψ − f) (1 − θε) − (Laϕ − f) (1 − θε)]v ≤ [f + (Laψ − f) ]v ˆΩ ˆΩ ˆΩ and + − − (Lauε)v = [f + (Laψ − f) (1 − θε) − (Laϕ − f) (1 − θε)]v ≥ [f − (Laϕ − f) ]v. ˆΩ ˆΩ ˆΩ Therefore,

− + [f ∧ (Laϕ)]v = [f − (Laϕ − f) ]v ≤ (Lauε)v ≤ [f + (Laψ − f) ]v = [f ∨ (Laψ)]v. ˆΩ ˆΩ ˆΩ ˆΩ ˆΩ Letting ε → 0, this holds for u. Since v is arbitrary, we conclude (4.5).

4.2.2 N-membranes problem

s We consider now the N-membranes problem, which consists of: To find u = (u1, u2, . . . , uN ) ∈ KN satisfying

N N X X E (u , v − u ) ≥ f i(v − u ), ∀(v , . . . , v ) ∈ s , (4.6) a i i i ˆ i i 1 N KN i=1 i=1 Ω

s s N where KN is the convex subset of [H0 (Ω)] defined by

s s N KN = {(v1, . . . , vN ) ∈ [H0 (Ω)] : v1 ≥ · · · ≥ vN a.e. in Ω} (4.7)

# and f i, . . . , f N ∈ L2 (Ω). As in the previous sections, the existence and uniqueness follows easily. Further- more, the following Lewy-Stampacchia type inequality also holds.

Theorem 4.7. The solution u = (u1, . . . uN ) of the N-membranes problem satisfies a.e. in Ω

1 1 N f ∧ Lau1 ≤ f ∨ · · · ∨ f 1 2 2 N f ∧ f ≤ Lau2 ≤ f ∨ · · · ∨ f . . 1 N−1 N−1 N f ∧ · · · ∧ f ≤ LauN−1 ≤ f ∨ f 1 N N f ∧ · · · ∧ f ≤ LauN ≤ f ,

2# N and Lau ∈ [L (Ω)] . Proof. Choosing (v, u , . . . , u ) ∈ s with v ∈ s , we see that u ∈ s solves (3.1) with f = f 1 and by 2 N KN Ku2 1 Ku2 Theorem 4.3 1 1 f ≤ Lau1 ≤ f ∨ Lau2 a.e. in Ω. Analogously, u ∈ s solves the two obstacles problem with f = f j, j = 2, 3,...,N − 1, and satisfies, j Kuj+1,uj−1 by (4.5), j j f ∧ Lauj−1 ≤ Lauj ≤ f ∨ Lauj+1 a.e. in Ω.

Finally, uN solves the one obstacle problem with an upper obstacle ϕ = uN−1, and so by the symmetric Lewy-Stampacchia estimates given in Theorem 4.3, we have

N N f ∧ LauN−1 ≤ LauN ≤ f a.e. in Ω.

The proof concludes by simple iteration (see Theorem 5.1 of [43]).

22 Remark 4.8. The solution can also approximated by the bounded penalization given in [7] by

s ε ε ε ε ε i s uε ∈ H0 (Ω) : hLaui , vii+ ζiθε(ui −ui+1)vi − ζi−1θε(ui−1 −ui )vi = (f +ζi −ζi−1)vi ∀vi ∈ H0 (Ω) ˆΩ ˆΩ ˆΩ for ( ) f 1 + ··· + f i ζ = max : i = 1,...,N , 0 i 1 i ζi = iζ0 − (f + ··· + f ) for i = 1,...,N

ε ε with θε(t) = 1 for t ≥ ε and u0 = +∞, uN+1 = −∞. ε s N As in [7], we then have the strong convergence of u → u in [H0 (Ω)] , thereby giving the Lewy- Stampacchia inequalities in Theorem 4.7.

4.3 Local regularity of solutions We make use of the Lewy-Stampacchia inequalities to show local regularity for the three types of nonlocal obstacle problems, but first it is useful to obtain the global boundedness of the solutions. Let s ∈ (0, 1). Suppose that

p d (a) f, Laψ ∈ L (Ω) for some p > 2s for the one obstacle problem,

p d (b) f ∧ Laϕ and f ∨ Laψ are in L (Ω) for some p > 2s for the two obstacles problem, or

i p d (c) f ∈ L (Ω) for i = 1,...,N for some p > 2s for the N-membranes problem. Theorem 4.9. Let u denote the solutions of the one obstacle problem (3.1), or the two obstacles problem (4.3), or u = ui for i = 1,...,N of the N-membranes problem (4.6), respectively, under the assumptions (a), p d (b) or (c) above. Then g = Lau ∈ L (Ω), with p > 2s and there exists a constant C, depending only on a∗, ∗ a , d, Ω, kuk s , kgk p and s, such that H0 (Ω) L (Ω)

kukL∞(Ω) ≤ C.

s Proof. Assume that Φ : R → R is a Lipschitz convex function such that Φ(0) = 0, then if u ∈ H0 (Ω), we have, by repeating the proof of Proposition 4 of [36],

0 s Ea(Φ(u), v) ≤ Ea(u, vΦ (u)) for v ≥ 0, v ∈ H0 (Ω), weakly in Ω.

We can then repeat the proof of Theorem 13 of [36] using the Moser technique to obtain the theorem.

Observe that in general our kernel does not satisfy the usual regularity of the kernel of the fractional Laplacian [45][47] or other commonly considered fractional kernels [12][27], since in general it does not satisfy the “symmetry” condition a(x, y) = a(x, −y) unless a is a constant multiple of the kernel of the fractional Laplacian. However, it will still be possible to obtain local H¨olderregularity on the solution with the properties of our kernel, if we assume it is symmetric, i.e. if it satisfies

a(x, y) = a(y, x). (4.8)

Then, we can make use of the symmetric form as given in (2.3), and apply the results of [36]. By the Lewy-Stampacchia inequalities, as long as the upper and lower bounds are in Lp(Ω) for some d p > 2s , we can make use of the Dirichlet form nature of the bilinear form, and obtain H¨olderregularity on the solutions on balls independently of the boundary conditions and of the regularity of ∂Ω. Then, by Theorem 1.6 of [26], since the bilinear form satisfies (2.2) and (2.4)–(2.5), in the symmetric case, we have the weak Harnack inequality. Furthermore, as in the classical de Giorgi-Nash-Moser theory, the weak Harnack inequality implies a decay of oscillation-result and local H¨olderregularity estimates for weak solutions.

23 Theorem 4.10 (Weak Harnack inequality). Let u denote the solutions of the one obstacle problem (3.1), or the two obstacles problem (4.3), or u = ui for i = 1,...,N of the N-membranes problem (4.6), respectively, under the assumptions (a), (b) or (c) above. Suppose the unit ball about the origin B1 is a subset of Ω, and a is symmetric. Then,

1 ! p0 p0 − inf u ≥ c u(x) dx − sup u (z)dµ (dz) −kL uk p , x a L (B15/16) B1/4 ˆ d B1/2 x∈B15/16 R \B1

for dµx(dz) a measure depending on a as defined in [26] and [34], and the positive constants p0 ∈ (0, 1) and ∗ c depend only on d, s, a∗, a . Theorem 4.11 (H¨olderregularity). Let u denote the solutions of the one obstacle problem (3.1), or the two obstacles problem (4.3), or u = ui for i = 1,...,N of the N-membranes problem (4.6), respectively, under the assumptions (a), (b) or (c) above. Suppose Bρ ⊂⊂ Ω is a ball of radius ρ and a is symmetric. Then, there exists cρ ≥ 0 and β ∈ (0, 1), independent of u, such that the following H¨olderestimate holds for almost every x, y ∈ Bρ/2: β   |u(x) − u(y)| ≤ c1|x − y| kukL∞(Ω) +kLaukLp(Ω) .

Proof. Since the Lewy-Stampacchia inequalities in Theorems 4.3, 4.6 and 4.7 hold a.e. in Bρ ⊂ Ω for Lau for the one obstacle, the two obstacles problem and the N-membranes problem respectively, and Lau = g in p d Bρ, therefore g lies in L (Ω) for p > 2s , and we have the result making use of Theorem 4.10 following the classical approach. Remark 4.12. (i) The local H¨oldercontinuity of the solutions of a two membranes problem was obtained for different operators with translation invariant kernels in [12], as well as the local C1,γ regularity in the case of the fractional Laplacian as in the case of the regular obstacles of [53]. (ii) The results in Theorems 4.10 and 4.11 can be generalized to arbitrary family of measures satisfying (2.2) and (2.4)–(2.5), as given in [14]. (iii) Since the L∞ bound also works in the non-symmetric case, we conjecture that the weak Harnack inequality and the H¨oldercontinuity are also true without the symmetry assumption, but this is an open problem.

s In the case where a corresponds to the kernel of the fractional Laplacian, La = (−∆) , we can use Corollary 1.15 of [50] and apply local elliptic regularity of weak solutions in fractional Sobolev spaces W r,p associated to Dirichlet fractional Laplacian problems as in Theorem 1.4 of [9].

# Theorem 4.13. Let u denote the solutions of the one obstacle problem (3.1) for f ∈ L2 (Ω) in the form

s s s s u ∈ Kψ(Ω) : D u · D (v − u) dx ≥ f(v − u) ∀v ∈ Kψ, ˆ d ˆ R Ω

or the corresponding two obstacles problem (4.3), or u = ui for i = 1,...,N of the corresponding N- membranes problem (4.6), respectively, under the assumptions (a), (b) or (c) above, with 2# ≤ p < ∞ s p 2s,p 1 and 0 < s < 1. Then, (−∆) u ∈ Lloc(Ω) and u ∈ Wloc (Ω). In particular, u ∈ C (Ω) if s > 1/2 and p > d/(2s − 1), by Theorem 7.57(c) of [4]. This theorem, which seems new, is an extension to nonlocal obstacle type problems of the well-known 2,p Wloc (Ω) regularity of solutions of the classical local obstacle problem corresponding to s = 1.

5 s-capacity and Lewy-Stampacchia Inequalities in H−s(Ω)

In this section, we extend the results on the Lewy-Stampacchia inequalities obtained in the previous section −s s to data in the dual space H (Ω). We first characterize the order dual of H0 (Ω), which is related to the theory of the s-capacity. This follows much of the results in the classical obstacle problem [55], [2] and [42].

24 In [3] and [29] the more general capacities are considered for general bilinear forms. Recently the fractional capacity for the Neumann problem was considered in [56]. In order to extend the results in Theorems 4.3, 4.6 and 4.7 to data in H−s(Ω), we may apply the general results of [38] for the one obstacle problem and [43] for two obstacles.

−s s 5.1 A characterization of the order dual H≺ (Ω) of H0 (Ω)

Associated with any Dirichlet form, there is a Choquet capacity. We denote by Cs the capacity associated s to the norm of H0 (Ω). For any compact set K ⊂ Ω, it is defined by

n 2 s o Cs(K) = inf kuk s : u ∈ H (Ω), u ≥ 1 a.e. in K . H0 (Ω) 0 For an arbitrary open set G ⊂ Ω,  Cs(G) = sup Cs(K) : K is a compact set in G .

s A function u ∈ H0 (Ω) is said to be quasi-continuous if for every ε > 0, there exists an open set G ⊂ Ω such that Cs(G) < ε and u|Ω\G is continuous. A property is said to hold quasi-everywhere (q.e.) if it holds except for a set of capacity zero. s It is well-known (by [3] Proposition 6.1.2 page 156 or [29] Theorem 2.1.3 page 71) that for every u ∈ H0 (Ω), there exists a unique (up to a set of capacity 0) quasi-continuous functionu ¯ : Ω → R such thatu ¯ = u a.e. on Ω. Therefore, we have the following theorem (see also Theorem 3.7 of [56]).

s Theorem 5.1. For every function u ∈ H0 (Ω), there exists a unique (up to q.e. equivalence) u¯ : Ω → R quasi-continuous function such that u =u ¯ a.e. in Ω.

s Thus, it makes sense to identify a function u ∈ H0 (Ω) with the class of quasi-continuous functions that are equivalent quasi-everywhere. Denote the space of such equivalent classes by Qs(Ω). Then, for every s element u ∈ H0 (Ω), there is an associatedu ¯ ∈ Qs(Ω). Define the space L2 (Ω) by Cs L2 (Ω) = {φ ∈ Q (Ω) : ∃u ∈ Hs(Ω) : u¯ ≥ |φ| q.e. in Ω} Cs s 0 and s RCs (φ) = inf{kuk s : u ∈ H (Ω), u¯ ≥ |φ| q.e.}, H0 (Ω) 0 which is a norm that makes L2 (Ω) a Banach space (see Proposition 1.2 of [6]). We want to show that the Cs dual space of L2 (Ω) can be identified with the order dual of Hs(Ω), i.e. Cs 0 [L2 (Ω)]0 = H−s(Ω) ∩ M(Ω) = H−s(Ω) = [H−s(Ω)]+ − [H−s(Ω)]+, Cs ≺ where M(Ω) is the set of bounded measures in Ω. Then we have the following result by Theorem 5.1 (corresponding to the classical case in [6] Proposition 1.7). Proposition 5.2. The injection of Hs(Ω) ∩ C (Ω) ,→ L2 (Ω) is dense. 0 c Cs Proof. This simply follows from Theorem 5.1, since u ≥ 0 a.e. on Ω implies u ≥ 0 q.e. on Ω.

s For K ⊂ Ω, recall that one says that u  0 on K (or u ≥ 0 on K in the sense of H0 (Ω)) if there exists a s sequence of Lipschitz functions uk → u in H0 (Ω) such that uk ≥ 0 on K. s Let K ⊂ Ω be any compact subset. Define the nonempty closed convex set of H0 (Ω) by s s KK = {v ∈ H0 (Ω) : v  1 on K}. Consider the following variational inequality

s s u ∈ KK : Ea(u, v − u) ≥ 0, ∀v ∈ KK . (5.1) This variational inequality has clearly a unique solution and consequently we can also extend to the s- fractional framework the following theorem which is due to Stampacchia [55] in the case s = 1.

25 Theorem 5.3 (Radon measure for the bilinear form Ea). For any compact K ⊂ Ω, the unique solution u of (5.1), which is called the (s, a)-capacitary potential of K, is such that

s u = 1 on K (in the sense of H0 (Ω))

µ = Lau ≥ 0 with supp(µ) ⊂ K. Moreover, for the non-negative Radon measure µ, one has

a Cs (K) = Ea(u, u) = dµ = µ(K) ˆΩ

and this number is called the (s, a)-capacity of K with respect to Ea(·, ·) (or to the operator La). Proof. The proof follows a similar approach to the classical case ([55] Theorem 3.9 or [42] Theorem 8.1). + s Taking v = u ∧ 1 = u − (u − 1) ∈ KK in (5.1), one has

+ 2 + + a∗ (u − 1) s ≤ Ea(u − 1, (u − 1) ) = Ea(u, (u − 1) ) ≤ 0 H0 (Ω)

s since the s-grad of a constant is zero. Hence u  1 in Ω. But u ∈ KK , so u  1 on K. Therefore, the first result u = 1 on K follows. s For the second result, set v = u + ϕ ∈ KK in (5.1) with an arbitrary ϕ ∈ D(Ω), ϕ ≥ 0. Then, by the Riesz-Schwartz theorem (see for instance [3] Theorem 1.1.3), there exists a non-negative Radon measure µ on Ω such that

hLau, ϕi = Ea(u, ϕ) = ϕ dµ, ∀ϕ ∈ D(Ω). ˆΩ s Moreover, for x ∈ Ω\K, there is a neighbourhood O ⊂ Ω\K of x so that u + ϕ ∈ KK for any ϕ ∈ D(O). Therefore, Ea(u, ϕ) = 0, ∀ϕ ∈ D(Ω\K)

which means µ = Lau = 0 in Ω\K. Therefore, supp(µ) ⊂ K and the third result follows immediately. We observe that when a corresponds to the kernel of the fractional Laplacian, the (s, a)-capacity corre- sponds to the s-capacity and the s-capacitary potential of a compact set K is the solution of the obstacle s problem (5.1) when the bilinear form is the inner product in H0 (Ω) and we have a simple comparison of the capacities in the following proposition.

Proposition 5.4. For any compact subset E ⊂ Ω,

∗2 a a a∗Cs(E) ≤ Cs (E) ≤ Cs(E). a∗ Proof. Let u be the (s, a)-capacitary potential of E, andu ¯ be the s-capacitary potential of E. Sinceu ¯  1 s on E, we can choose v =u ¯ ∈ KE in (5.1) to get

a Cs (E) = Ea(u, u) ≤Ea(u, u¯) 1 1   2   2 ≤a∗ |Dsu|2 |Dsu¯|2 ˆ d ˆ d R R a a∗2 ≤ ∗ |Dsu|2 + |Dsu¯|2 2 ˆ d 2a ˆ d R ∗ R 1 a∗2 ≤ Ea(u, u) + Cs(E) 2 2a∗ ∗2 1 a a = Cs (E) + Cs(E) 2 2a∗

26 s by Cauchy-Schwarz inequality and the coercivity of a. Similarly, we can choose v = u ∈ KE for (5.1) with a being the identity for Cs(E) to get, using the coercivity of a,

Cs(E) = Ea(¯u, u¯) ≤Ea(¯u, u) 1 1   2   2 ≤ |Dsu¯|2 |Dsu|2 ˆ d ˆ d R R 1 1 ≤ |Dsu¯|2 + |Dsu|2 2 ˆ d 2 ˆ d R R 1 1 ≤ Cs(E) + Ea(u, u) 2 2a∗ 1 1 a = Cs(E) + Cs (E). 2 2a∗

Using this definition of the Radon measure, we recall that two quasi-continuous functions which are equal (or, ≤) µ-a.e. on an open subset of Rd are also equal (or, ≤) q.e. on that set (see [29] Lemma 2.1.4). s Recall that a Radon measure µ is said to be of finite energy relatively to H0 (Ω) if its restriction to s s H0 (Ω) ∩ Cc(Ω) is continuous for the topology of H0 (Ω), by means of

s hµ, vi = v dµ, ∀v ∈ H0 (Ω) ∩ Cc(Ω). ˆΩ Such a finite energy measure can in fact be defined for any Dirichlet form E (see [29] Section 2.2 and Example + s s 2.2.1 pages 87–91). We denote by E (H0 (Ω)) the cone of positive finite energy measures relative to H0 (Ω). −s Then µ is of finite energy if and only if there exists wµ ∈ H (Ω) such that

s hwµ, vi = v dµ ∀v ∈ H0 (Ω) ∩ Cc(Ω), ˆΩ

+ s −s + −s s 0 and E (H0 (Ω)) can be identified with [H (Ω)] , the positive cone of H (Ω) = [H0 (Ω)] , by the mapping + s s s µ 7→ wµ. Moreover, whenever µ ∈ E (H0 (Ω)), the mapping u ∈ H0 (Ω) 7→ u¯ is continuous from H0 (Ω) in 1 s s L (µ) and whenever u ∈ H0 (Ω), Ω u¯ dµ = hwµ, vi. Note that in the particular case of the space H0 (Ω), the s 1 mapping u ∈ H0 (Ω) 7→ u¯ ∈ L (µ)´ is compact; this follows from the fact that Ω |u¯n| dµ = hwµ, |un|i and that s s if un * 0 in H0 (Ω) then |un| * 0 in H0 (Ω). ´ Extending these results to L2 (Ω), we have the following result. Cs Proposition 5.5. Let µ ∈ E+(Hs(Ω)). Then L2 (Ω) ⊂ L1(µ) and this inclusion is continuous. 0 Cs Proof. Let u ∈ L2 (Ω). There exists v ∈ Hs(Ω) such thatv ¯ ≥ |u| a.e., and therefore µ-q.e.. Sincev ¯ ∈ L1(µ), Cs 0 u ∈ L1(µ). Let (u ) be a sequence in L2 (Ω) such that R (u ) → 0. Then there exists (v ) ∈ Hs(Ω) such that n Cs Cs n n 0 v¯n ≥ |un| q.e., and therefore µ-q.e., and kvnk s → 0. As a result, |un| dµ ≤ v¯n dµ = hwµ, vni ≤ H0 (Ω) Ω Ω 1 ´ ´ kwµkH−s(Ω)kvnk s → 0. Therefore un → 0 in L (µ). H0 (Ω) Having these results, we can now identify the dual space of L2 (Ω) with the order dual of Hs(Ω), as Cs 0 given in the following theorem.

Theorem 5.6 (Characterization of Order Dual). The dual of L2 (Ω) is the space of finite energy measures Cs + s + s −s s E (H0 (Ω)) − E (H0 (Ω)), that is identified with the order dual H≺ (Ω) of H0 (Ω). More precisely, L ∈ [L2 (Ω)]0 if and only if there is a unique µ such as |µ| ∈ E+(Hs(Ω)) and L(φ) = φ dµ for all φ ∈ L2 (Ω). Cs 0 Ω Cs 2 0 In addition, the norm of L in [L (Ω)] is such that kLk = kw k −s . ´ Cs |µ| H (Ω) Proof. According to Proposition 5.2, C (Ω) is dense in L2 (Ω) and moreover this injection is continuous; c Cs therefore the dual of L2 (Ω) is a space of measures. Cs

27 Let µ be a Radon measure such that |µ| ∈ E+(Hs(Ω)). For any φ ∈ L2 (Ω) (φ is then µ integrable by 0 Cs s Proposition 5.5), set L(φ) = Ω φ dµ. For any v ∈ H0 such thatv ¯ ≥ |φ| quasi-everywhere, so µ-a.e., we have ´

|L(φ)| = φ dµ ≤ |φ|d|µ| ≤ vd¯ |µ| = (w|µ|, v) ≤ kw|µ|kH−s(Ω)kvnkHs(Ω) , ˆΩ ˆΩ 0

2 0 so |L(φ)| ≤ kw kR (φ). Therefore L ∈ [L (Ω)] and kLk ≤ kw k −s . |µ| Cs Cs |µ| H (Ω) Conversely, suppose L ∈ [L2 (Ω)]0. Let K be a compact subset of Ω and ψ ∈ D(Ω) such that ψ ≥ 1 in Cs K. Whenever φ ∈ Cs(K) such that kφk∞ ≤ 1, we have ψ ≥ |φ| in Ω, therefore

|L(φ)| ≤kLk · RCs (ψ).

We deduce that there exists a Radon measure µ in Ω such that

∀φ ∈ Cc(Ω),L(φ) = φ dµ. ˆΩ

+ s s In addition |µ| ∈ E (H0 ) because, whenever u ∈ H0 (Ω) ∩ Cc(Ω),

ud|µ| ≤ |u|d|µ| ˆΩ ˆΩ

= sup{L(φ) : φ ∈ Cc(Ω), |φ| ≤ |u|}

≤ sup{kLk · RCs (φ) : |φ| ≤ |u|}

≤kLk · RCs (u)

≤kLk ·kuk s H0 (Ω) and also kw|µ|kH−s(Ω) ≤kLk. Finally, it follows, from the density of C (Ω) in L2 (Ω) and Proposition 5.5, that c Cs

∀φ ∈ L2 (Ω),L(φ) = φ dµ. Cs ˆΩ

−s 5.2 Lewy-Stampacchia inequalities in H≺ (Ω) −s For completeness, we state the Lewy-Stampacchia inequalities in the dual space H≺ (Ω).

Theorem 5.7. The unique solution u to the obstacle problem (3.1) with compatible obstacle ψ ∈ Hs(Rd) −s and F, Laψ ∈ H≺ (Ω), satisfies −s F ≤ Lau ≤ F ∨ Laψ in H≺ (Ω). (5.2)

Proof. Since La is strictly T-monotone, this is a direct consequence of the abstract Lewy-Stampacchia inequality obtained by Mosco in [38] (see also Theorem 2.1 of [42]). We next consider the generalizations to the two obstacles problem and to the N-membranes problem. Similarly, as a direct consequence of Theorem 4.2 of [43], we may also state the Lewy-Stampacchia inequality for the two obstacles problem. Theorem 5.8. The solution u of the two obstacles problem

s s u ∈ Kϕ,ψ : Ea(u, v − u) ≥ hF, v − ui ∀v ∈ Kϕ,ψ

s −s −s with Kϕ,ψ given by (4.4) with data F ∈ H≺ (Ω) and Laψ, Laϕ ∈ H≺ (Ω) satisfies

−s F ∧ Laϕ ≤ Lau ≤ F ∨ Laψ in H≺ (Ω). (5.3)

28 Then, applying the general Lewy-Stampacchia inequalities for the one obstacle and for the two obstacles problem iteratively in the previous theorem as in the proof of Theorem 4.7, we obtain Theorem 5.9. The solution u of the N-membranes problem

N N s X X i s u ∈ KN : Ea(ui, vi − ui) ≥ hF , vi − uii ∀(v1, . . . vN ) ∈ KN i=1 i=1

s 1 N i −s with KN given by (4.7) with data F = (F ,...,F ) for F ∈ H≺ (Ω) satisfies 1 1 N F ∧ Lau1 ≤ F ∨ · · · ∨ F 1 2 2 N F ∧ F ≤ Lau2 ≤ F ∨ · · · ∨ F . . 1 N−1 N−1 N F ∧ · · · ∧ F ≤ LauN−1 ≤ F ∨ F 1 N N F ∧ · · · ∧ F ≤ LauN ≤ F

−s in H≺ (Ω). Remark 5.10. In the symmetric case, the Lewy-Stampacchia inequalities also follow from the general results of [30]. The application to Theorem 5.9 for the N-membranes problem is new.

5.3 The Ea obstacle problem and the s-capacity As a simple application of s-capacity, we consider the corresponding nonlocal obstacle problem extending some results of [55] and [2] (see also [42]). In this section we start by the following comparison property for the (s, a)-capacity which proof is exactly as in Theorem 3.10 of [55], which states that in the case when the kernel a is symmetric the (s, a)-capacity is an increasing set function.

Proposition 5.11. For any compact subsets E1 ⊂ E2 ⊂ Ω,  2 a M a Cs (E1) ≤ 1 + Cs (E2), a∗

1 where M = sup (Ea(u, v) − Ea(v, u)) for u, v such that kuk s =kvk s = 1. 2 H0 (Ω) H0 (Ω) Theorem 5.12. Let ψ be an arbitrary function in L2 (Ω). Suppose that the closed convex set ¯ s is such Cs Kψ that ¯ s s Kψ = {v ∈ H0 (Ω) : v¯ ≥ ψ q.e. in Ω}= 6 ∅. Then there is a unique solution to ¯ s ¯ s u ∈ Kψ : Ea(u, v − u) ≥ 0, ∀v ∈ Kψ, (5.4) which is non-negative and such that

∗ + kukHs(Ω) ≤ (a /a∗) ψ L2 (Ω) ; (5.5) 0 Cs

+ there is a unique measure µ = Lau ≥ 0, concentrated on the coincidence set {u = ψ} = {u = ψ }, verifying

s Ea(u, v) = v¯ dµ, ∀v ∈ H0 (Ω), (5.6) ˆΩ and ∗2 ! a +  a 1/2 µ(K) ≤ ψ 2 Cs (K) , ∀K ⊂⊂ Ω, (5.7) 3/2 LC (Ω) a∗ s in particular µ does not charge on sets of capacity zero.

29 Proof. By the maximum principle Proposition 3.4(ii), taking v = u+u−, the solution is non-negative. Hence, ¯ s ¯ s the variational inequality (5.4) is equivalent to solving the variational inequality with Kψ replaced by Kψ+ . Since ψ+ ∈ L2 (Ω), by definition of L2 (Ω), ¯ s 6= ∅ and we can apply the Stampacchia theorem to obtain Cs Cs Kψ+ a unique non-negative solution. Since Ea(u, v − u) ≥ 0,

2 ∗ a∗kuk s ≤ Ea(u, u) ≤ Ea(u, v) ≤ a kuk s kvk s , H0 (Ω) H0 (Ω) H0 (Ω) we have ∗ ¯ s kuk s ≤ (a /a∗)kvk s , ∀v ∈ + , H0 (Ω) H0 (Ω) Kψ which, by the definition of the L2 (Ω) norm of ψ+, gives (5.5). Cs The existence of Radon measure for (5.6) follows exactly as in Theorem 5.3. Finally, recalling the definitions, it is sufficient to prove (5.7) for any compact subset K ⊂ Ω. But this follows from

∗ ∗2 + s µ(K) ≤ v¯ dµ = Ea(u, v) ≤ a kukHs(Ω)kvkHs(Ω) ≤ (a /a∗) ψ L2 (Ω)kvkHs(Ω) , ∀v ∈ KK . ˆΩ 0 0 Cs 0 Now, recall from Proposition 5.4 that we have a 2 C (K) ≥ a∗Cs(K) = a∗ inf kvk s s s H0 (Ω) v∈KK thereby obtaining (5.7).

Corollary 5.13. If u and uˆ are the solutions to (5.4) with non-negative compatible obstacles ψ and ψˆ in L2 (Ω) respectively, then Cs ˆ 1/2 ku − uˆkHs(Ω) ≤ kkψ − ψkL2 (Ω), 0 Cs where 1/2 ∗ h ˆ i k = (a /a∗) kψkL2 (Ω) + kψkL2 (Ω) . Cs Cs ˆ Proof. Since supp(µ) ⊂ {u = ψ} and supp(ˆµ) ⊂ {uˆ = ψ} (where µ = Lau andµ ˆ = Lauˆ), for an arbitrary v ∈ K¯ s , we have |ψ−ψˆ| 2 a∗ku − uˆk s ≤ Ea(u − u,ˆ u − uˆ) H0 (Ω)

= Ea(u, u − uˆ) − Ea(ˆu, u − uˆ)

= (u − uˆ) dµ − (u − uˆ)dµˆ by settingv ¯ = u − uˆ for µ, µˆ in (5.6) ˆΩ ˆΩ ≤ (ψ − ψˆ) dµ − (ψ − ψˆ)dµˆ ˆΩ ˆΩ ≤ |ψ − ψˆ|d(µ +µ ˆ) ˆΩ ≤ vd¯ (µ +µ ˆ) since v ∈ K¯ s |ψ−ψˆ| ˆΩ = v¯ dµ + vd¯ µˆ ˆΩ ˆΩ = Ea(u, v) + Ea(ˆu, v) ∗ h i ≤ a kuk s +kuˆk s kvk s H0 (Ω) H0 (Ω) H0 (Ω) ∗2 a h ˆ i ≤ kψkL2 (Ω) + kψkL2 (Ω) kvkHs(Ω) by (5.5) a∗ Cs Cs 0 ∗2 a h ˆ i ˆ ≤ kψkL2 (Ω) + kψkL2 (Ω) kψ − ψkL2 (Ω) a∗ Cs Cs Cs since v is arbitrary in K¯ s , by the definition of the norm of |ψ − ψˆ| in L2 (Ω). |ψ−ψˆ| Cs

30 Remark 5.14. Further properties on the s-capacity and the regularity of the solution to the s-obstacle problem are an interesting topic to be developed. For instance, as in the classical local case s = 1 [2], it ¯ s would be interesting to show that ψ is compatible, i.e. Kψ 6= ∅, if and only if

∞ + 2 Cs({|ψ | > t})dt < ∞. ˆ0 Acknowledgements. C. Lo acknowledges the FCT PhD fellowship in the framework of the LisMath doctoral programme at the University of Lisbon. The research of J. F. Rodrigues was partially done under the framework of the Project PTDC/MATPUR/28686/2017 at CMAFcIO/ULisboa. Also, many thanks to Alan Aw, Pedro Campos, Lisa Santos, Shuang Wu, and an anonymous referee, for providing useful comments.

References

[1] Nicola Abatangelo and Xavier Ros-Oton. Obstacle problems for integro-differential operators: higher regularity of free boundaries. Adv. Math., 360:106931, 61, 2020. [2] David R. Adams. Capacity and the obstacle problem. Appl. Math. Optim., 8(1):39–57, 1982. [3] David R. Adams and Lars Inge Hedberg. Function spaces and , volume 314 of Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences]. Springer-Verlag, Berlin, 1996. [4] Robert A. Adams. Sobolev spaces. Academic Press [A subsidiary of Harcourt Brace Jovanovich, Pub- lishers], New York-London, 1975. Pure and Applied , Vol. 65. [5] Harbir Antil and Carlos N. Rautenberg. Fractional elliptic quasi-variational inequalities: theory and numerics. Interfaces Free Bound., 20(1):1–24, 2018. [6] H´edyAttouch and Colette Picard. In´equationsvariationnelles avec obstacles et espaces fonctionnels en th´eoriedu potentiel. Applicable Anal., 12(4):287–306, 1981. [7] Assis Azevedo, Jos´e-Francisco Rodrigues, and Lisa Santos. The N-membranes problem for quasilinear degenerate systems. Interfaces Free Bound., 7(3):319–337, 2005. [8] Jos´eC. Bellido, Javier Cueto, and Carlos Mora-Corral. Γ-convergence of polyconvex functionals in- volving s-fractional gradients to their local counterparts. Calc. Var. Partial Differential Equations, 60(1):Paper No. 7, 2021. [9] Umberto Biccari, Mahamadi Warma, and Enrique Zuazua. Local elliptic regularity for the Dirichlet fractional Laplacian. Adv. Nonlinear Stud., 17(2):387–409, 2017. [10] Andrea Bonito, Wenyu Lei, and Abner J. Salgado. Finite element approximation of an obstacle problem for a class of integro-differential operators. ESAIM Math. Model. Numer. Anal., 54(1):229–253, 2020. [11] Xavier Cabr´eand Yannick Sire. Nonlinear equations for fractional Laplacians, I: Regularity, maximum principles, and Hamiltonian estimates. Ann. Inst. H. Poincar´eAnal. Non Lin´eaire, 31(1):23–53, 2014. [12] Luis Caffarelli, Daniela De Silva, and Ovidiu V. Savin. The two membranes problem for different operators. Ann. Inst. H. Poincar´eAnal. Non Lin´eaire, 34(4):899–932, 2017. [13] Luis A. Caffarelli, Sandro Salsa, and Luis Silvestre. Regularity estimates for the solution and the free boundary of the obstacle problem for the fractional Laplacian. Invent. Math., 171(2):425–461, 2008. [14] Jamil Chaker and Moritz Kassmann. Nonlocal operators with singular anisotropic kernels. Comm. Partial Differential Equations, 45(1):1–31, 2020. [15] Giovanni E. Comi and Giorgio Stefani. A distributional approach to fractional Sobolev spaces and fractional variation: asymptotics i. arXiv: 1910.13419 [math.FA], 2019.

31 [16] Giovanni E. Comi and Giorgio Stefani. A distributional approach to fractional Sobolev spaces and fractional variation: Existence of blow-up. Journal of Functional Analysis, 277(10):3373 – 3435, 2019. [17] Donatella Danielli, Arshak Petrosyan, and Camelia A. Pop. Obstacle problems for nonlocal operators. In New developments in the analysis of nonlocal operators, volume 723 of Contemp. Math., pages 191–214. Amer. Math. Soc., Providence, RI, 2019.

[18] Donatella Danielli and Sandro Salsa. Obstacle problems involving the fractional Laplacian. In Recent developments in nonlocal theory, pages 81–164. De Gruyter, Berlin, 2018. [19] Marta D’Elia, Mamikon Gulian, Hayley Olson, and George Em Karniadakis. A unified theory of fractional, nonlocal, and weighted nonlocal vector calculus. arXiv preprint arXiv:2005.07686, 5 2020.

[20] Marta D’Elia and Mamikon Gulian. Analysis of anisotropic nonlocal diffusion models: Well-posedness of fractional problems for anomalous transport. arXiv preprint arXiv:2101.04289, 1 2021. [21] Fran¸coiseDemengel and Gilbert Demengel. Functional spaces for the theory of elliptic partial differential equations. Universitext. Springer, London; EDP Sciences, Les Ulis, 2012. Translated from the 2007 French original by Reinie Ern´e.

[22] Jean-Dominique Deuschel and Takashi Kumagai. Markov chain approximations to nonsymmetric diffu- sions with bounded coefficients. Comm. Pure Appl. Math., 66(6):821–866, 2013. [23] Eleonora Di Nezza, Giampiero Palatucci, and Enrico Valdinoci Hitchhiker’s guide to the fractional Sobolev spaces. Bull. Sci. Math. 136(5):521–573, 2012.

[24] Qiang Du, Max Gunzburger, Richard B. Lehoucq, and Kun Zhou. Analysis and approximation of nonlocal diffusion problems with volume constraints. SIAM Rev., 54(4):667–696, 2012. [25] Qiang Du, Max Gunzburger, Richard B. Lehoucq, and Kun Zhou. A nonlocal vector calculus, nonlocal volume-constrained problems, and nonlocal balance laws. Math. Models Methods Appl. Sci., 23(3):493– 540, 2013.

[26] Bartlomiej Dyda and Moritz Kassmann. Regularity estimates for elliptic nonlocal operators. Anal. PDE, 13(2):317–370, 2020. [27] Mouhamed Moustapha Fall. Regularity results for nonlocal equations and applications. Calc. Var. Partial Differential Equations, 59(5):Paper No. 181, 53, 2020.

[28] Matthieu Felsinger, Moritz Kassmann, and Paul Voigt. The Dirichlet problem for nonlocal operators. Math. Z., 279(3-4):779–809, 2015. [29] Masatoshi Fukushima, Yoichi Oshima, and Masayoshi Takeda. Dirichlet forms and symmetric Markov processes, volume 19 of De Gruyter Studies in Mathematics. Walter de Gruyter & Co., Berlin, extended edition, 2011.

[30] Nicola Gigli and Sunra Mosconi. The abstract Lewy-Stampacchia inequality and applications. J. Math. Pures Appl. (9), 104(2):258–275, 2015. [31] Max Gunzburger and Richard B. Lehoucq. A nonlocal vector calculus with application to nonlocal boundary value problems. Multiscale Model. Simul., 8(5):1581–1598, 2010.

[32] Moritz Kassmann. The theory of De Giorgi for non-local operators. C. R. Math. Acad. Sci. Paris, 345(11):621–624, 2007. [33] Moritz Kassmann. A priori estimates for integro-differential operators with measurable kernels. Calc. Var. Partial Differential Equations, 34(1):1–21, 2009. [34] Moritz Kassmann. A new formulation of Harnack’s inequality for nonlocal operators. C. R. Math. Acad. Sci. Paris, 349(11-12):637–640, 2011.

32 [35] David Kinderlehrer and . An introduction to variational inequalities and their appli- cations, volume 31 of Classics in Applied Mathematics. Society for Industrial and Applied Mathematics (SIAM), Philadelphia, PA, 2000. Reprint of the 1980 original. [36] Tommaso Leonori, Ireneo Peral, Ana Primo, and Fernando Soria. Basic estimates for solutions of a class of nonlocal elliptic and parabolic equations. Discrete Contin. Dyn. Syst., 35(12):6031–6068, 2015.

[37] Zhi Ming Ma and Michael R¨ockner. Introduction to the theory of (nonsymmetric) Dirichlet forms. Universitext. Springer-Verlag, Berlin, 1992. [38] Umberto Mosco. Implicit variational problems and quasi variational inequalities. In Nonlinear operators and the (Summer School, Univ. Libre Bruxelles, Brussels, 1975), pages 83–156. Lecture Notes in Math., Vol. 543. 1976. [39] Roberta Musina and Alexander I. Nazarov. Variational inequalities for the spectral fractional Laplacian. Comput. Math. Math. Phys., 57(3):373–386, 2017. [40] Roberta Musina, Alexander I. Nazarov, and Konijeti Sreenadh. Variational inequalities for the fractional Laplacian. Potential Anal., 46(3):485–498, 2017.

[41] Ricardo H. Nochetto, Enrique Ot´arola,and Abner J. Salgado. Convergence rates for the classical, thin and fractional elliptic obstacle problems. Philos. Trans. Roy. Soc. A, 373(2050):20140449, 14, 2015. [42] Jos´e-Francisco Rodrigues. Obstacle problems in mathematical physics, volume 134 of North-Holland Mathematics Studies. North-Holland Publishing Co., Amsterdam, 1987. Notas de Matem´atica[Mathe- matical Notes], 114. [43] Jos´eFrancisco Rodrigues and Rafayel Teymurazyan. On the two obstacles problem in Orlicz-Sobolev spaces and applications. Complex Var. Elliptic Equ., 56(7-9):769–787, 2011. [44] Jos´eFrancisco Rodrigues and Lisa Santos. On nonlocal variational and quasi-variational inequalities with fractional gradient. Applied Mathematics & Optimization, 80, no. 3:835 – 852, 2019 (see Correction in https://arxiv.org/pdf/1903.02646.pdf). [45] Xavier Ros-Oton. Nonlocal elliptic equations in bounded domains: a survey. Publ. Mat., 60(1):3–26, 2016. [46] Xavier Ros-Oton. Obstacle problems and free boundaries: an overview. SeMA J., 75(3):399–419, 2018.

[47] Xavier Ros-Oton and Joaquim Serra. The Dirichlet problem for the fractional Laplacian: regularity up to the boundary. J. Math. Pures Appl. (9), 101(3):275–302, 2014. [48] Sandro Salsa. Optimal regularity in lower dimensional obstacle problems. In Subelliptic PDE’s and applications to geometry and finance, volume 6 of Lect. Notes Semin. Interdiscip. Mat., pages 217–226. Semin. Interdiscip. Mat. (S.I.M.), Potenza, 2007.

[49] Raffaella Servadei and Enrico Valdinoci. Lewy-Stampacchia type estimates for variational inequalities driven by (non)local operators. Rev. Mat. Iberoam., 29(3):1091–1126, 2013. [50] Tien-Tsan Shieh and Daniel Spector. On a new class of fractional partial differential equations. Advances in Calculus of Variations, 8:321 – 366, 2014.

[51] Tien-Tsan Shieh and Daniel Spector. On a new class of fractional partial differential equations ii. Advances in Calculus of Variations, 11:289 – 307, 2017. [52] Miroslav Silhavy. Fractional vector analysis based on invariance requirements (critique of coordinate approaches). Continuum Mechanics and Thermodynamics, 32, Issue 1:207 – 288, 2020.

[53] Luis Silvestre. Regularity of the obstacle problem for a fractional power of the Laplace operator. Comm. Pure Appl. Math., 60(1):67–112, 2007.

33 [54] Luis Enrique Silvestre. Regularity of the obstacle problem for a fractional power of the Laplace operator. ProQuest LLC, Ann Arbor, MI, 2005. Thesis (Ph.D.)–The University of Texas at Austin. [55] Guido Stampacchia. Le probl`emede Dirichlet pour les ´equationselliptiques du second ordre `acoefficients discontinus. Ann. Inst. Fourier (Grenoble), 15(fasc. 1):189–258, 1965.

[56] Mahamadi Warma. The fractional relative capacity and the fractional Laplacian with Neumann and Robin boundary conditions on open sets. Potential Anal., 42(2):499–547, 2015.

34