arxiv: v3 [math.ap] 20 Jun 2017

Size: px
Start display at page:

Download "arxiv: v3 [math.ap] 20 Jun 2017"

Transcription

1 ON A LONG RANGE SEGREGATION MODEL L. CAFFARELLI, S. PATRIZI, AND V. QUITALO arxiv: v3 [math.ap] 20 Jun 2017 Abstract. In this work we study the properties of segregation processes modeled by a family of equations L(u i)(x) = u i(x) F i(u 1,..., u K)(x) i = 1,..., K where F i(u 1,..., u K)(x) is a non-local factor that takes into consideration the values of the functions u j s in a full neighborhood of x. We consider as a model problem u ε i (x) = 1 ε 2 uε i (x) i j H(u ε j)(x) where ε is a small parameter and H(u ε j)(x) is for instance H(u ε j)(x) = u ε j(y) dy or H(u ε j)(x) = sup y B 1 (x) u ε j(y). Here the set B 1(x) is the unit ball centered at x with respect to a smooth, uniformly convex norm ρ of R n. Heuristically, this will force the populations to stay at ρ-distance 1, one from each other, as ε 0. B 1 (x) 1. Introduction Segregation phenomena occur in many areas of mathematics and science: from equipartition problems in geometry, to social and biological processes (cells, bacteria, ants, mammals), to finance (sellers and buyers). There is a large body of literature in connection to our work and we would like to refer to [4, 5, 8 21, 26 29, 31 33] and the references therein. We particularly would like to point out the articles [15,26,28,29,31] where spatial separation due to competition for resources is discussed among ant nests, mussels and sessile animals. They study a family of models arising from different applications whose main two ingredients are: in the absence of competition species follow a propagation equation involving diffusion, Date: June 21, Mathematics Subject Classification. Primary: 35J60; Secondary: 35R35, 35B65, 35Q92. Key words and phrases. Regularity for viscosity solutions, Segregation of populations. S. Patrizi was supported by the ERC grant EPSILON Elliptic Pde s and Symmetry of Interfaces and Layers for Odd Nonlinearities. 1

2 transport, birth-death, etc, but when two species overlap, their growth is mutually inhibited by competition, consumption of resources, etc. The simplest form of such models consists, for species σ i with spatial density u i, on a system of equations L(u i ) = u i F i (u 1,..., u K ). The operator L quantifies diffusion, transport, etc, while the term u i F i does attrition of u i from competition with the remaining species. In these models, the interaction is punctual, i.e. u i (x) interacts with the remaining densities also at position x. There are many processes, though where the growth of σ i at x is inhibited by the populations σ j in a full area surrounding x. The purpose of this work is a first attempt to study the properties of such a segregation process. Basically, we consider a family of equations, L(u i )(x) = u i (x) F i (u 1,..., u K )(x) where F i (u 1,..., u K )(x) is now a non-local factor that takes into consideration the values of u j in a full neighborhood of x. Given the previous discussion a possible model problem would be the system u ε i (x) = 1 ε 2 uε i (x) i j H(u ε j)(x), i = 1,..., K where ε is a small parameter and H(u ε j )(x) is a non-local operator, for instance H(u ε j)(x) = u ε j(y) dy or H(u ε j)(x) = B 1 (x) sup u ε j(y). y B 1 (x) To study the limit configuration when the competition for resources is very high, we consider the limit when ε tends to 0. Heuristically, the non-local term forces the populations to stay at distance 1, one from each other. As an example, as we will prove, in the case of two populations in dimension two, we will have strips of length precisely one between the regions 2

3 where the populations live. At edge points, that we will define as singular points, the angles of the asymptotic cones have to be the same, see Figure 1. Here S i = Si 1 S2 i, i = 1, 2, represents the region where the the population σ i with density u i exists. Moreover, the ratio between the normal derivatives at regular points across the free boundary, depends on the ratio of the respective curvature κ. For example, if Z 1 S 1 1 and Z 2 S 1 2, Z 1 and Z 2 are not edge points, and d(z 1, Z 2 ) = 1 then u 1 ν(z 1 ) u 2 ν(z 2 ) = κ(z 1) κ(z 2 ) if κ(z 2 ) 0, and u 1 ν(z 1 ) = u 2 ν(z 2 ) if κ(z 2 ) = 0. S 2 1 S 1 2 S 2 2 Z 2 Z 1 S 1 1 Figure 1. Example of a limit configuration for K = 2, n = 2 We will consider instead of the unit ball in the Euclidean norm B 1 (x), the translation at x of a general smooth set B that is also uniformly convex, bounded and symmetric with respect to the origin. The set B defines a smooth, uniformly convex norm ρ in R n. Let us note that there is some similarity also with the Lasry-Lions model of price formation (see [6, 25]) where selling and buying prices are separated by a gap due to transaction cost. 3

4 2. Notation and statement of the problem Let B be an open bounded domain of R n, convex, symmetric with respect to the origin and with smooth boundary. Then B can be represented as the unit ball of a norm ρ : R n R, ρ C (R n \ {0}), called the defining function of B, i.e., B = {x R n ρ(x) < 1}. We assume that B is uniformly convex, i.e., there exists 0 < a A such that in R n \ {0} (2.1) ai n D 2 ( 1 2 ρ2 ) AI n, where I n is the n n identity matrix. In what follows we denote B r := {y R n ρ(y) < r}, B r (x) := {y R n ρ(x y) < r}. So through the paper we will always refer to the Euclidean ball as B and to the ρ-ball as B. For a given closed set K, let d ρ (, K) = inf ρ( y) y K be the distance function from K associated to ρ. Then there exist c 1, c 2 > 0 such that (2.2) c 1 d(, K) d ρ (, K) c 2 d(, K), where d(, K) is the distance function associated to the Euclidian norm of R n. Let Ω R n be a bounded Lipschitz domain. We will denote by ( Ω) 1 the ρ-strip of size 1 around Ω in the complement of Ω defined by ( Ω) 1 := {x Ω c : d ρ (x, Ω) 1}. For i = 1,..., K, let f i be non-negative functions defined on ( Ω) 1 with supports at ρ-distance equal or greater than 1, one from each other: (2.3) d ρ (supp f i, supp f j ) 1, for i j. 4

5 We will consider the following system of equations: for i = 1,..., K u ε i (x) = 1 ε (2.4) 2 uε i (x) H(u ε j)(x) in Ω, j i u ε i = f i on ( Ω) 1. The functional H(u j )(x) depends only on the restriction of u j to B 1 (x). We will consider, for simplicity, (2.5) H(w)(x) = B 1 (x) w p (y)ϕ ( ρ(x y) ) dy, 1 p < or (2.6) H(w) ( x ) = sup w B 1 (x) with ϕ a strictly positive smooth function of ρ, with at most polynomial decay at B 1 : (2.7) ϕ(ρ) C(1 ρ) q, q 0. In rest of the paper, when we refer to consider u ε 1,..., uε K, viscosity solutions of the problem (2.4), we mean that u ε 1,..., uε K are continuous functions that satisfy in the viscosity sense the system of equations (2.4). Moreover, we make the following assumptions: for i = 1,..., K, ε > 0, Ω is a bounded Lipschitz domain of R n, f i : ( Ω) 1 R, f i 0, f i 0, f i is Hölder continuous, (2.8) c > 0 s. t. x Ω supp f i, B r (x) supp f i c B r (x), (2.3) holds true, H is either of the form (2.5) or (2.6) and (2.7) holds. 3. Main results For the reader s convenience we present our main results below. Assume that (2.8) holds true, then: Existence (Theorem 4.1): There exist continuous functions u ε 1,..., uε K, depending on the parameter ε, viscosity solutions of the problem (2.4). 5

6 Limit problem (Corollary 5.6): There exists a subsequence ( u) εm converging locally uniformly, as ε 0, to a function u = (u 1,..., u K ), satisfying the following properties: i) the u i s are locally Lipschitz continuous in Ω and have supports at distance at least 1, one from each other, i.e. u i 0 in the set {x Ω d ρ (x, supp u j ) 1} for any j i. ii) u i = 0 when u i > 0. Semiconvexity of the free boundary (Corollary 6.2): If x 0 {u i > 0} there is an exterior tangent ρ-ball of radius 1 at x 0. The supports of u i are sets of finite perimeter (Corollary 6.5): The set {u i > 0} has finite perimeter. Sharp characterization of the interfaces (Theorem 7.1): Under the additional assumption that p = 1 in (2.5), the supports of the limit functions are at distance exactly 1, one from each other, i.e, if x 0 {u i > 0} Ω, then there exists j i such that B 1 (x 0 ) {u j > 0}. Classification of singular points in dimension 2 (Lemma 8.9, Theorem 8.10, Corollary 8.11, Corollary 8.12): For n = 2, under the additional assumption that p = 1 in (2.5), for i j, let x 0 {u i > 0} Ω and y 0 {u j > 0} Ω be points such that {u i > 0} has an angle θ i at x 0, {u j > 0} has an angle θ j at y 0 and ρ(x 0 y 0 ) = 1. Then we have θ i = θ j. If x 0 {u i > 0} Ω and y 0 {u j > 0} Ω, then θ i θ j. 6

7 Moreover, singular points, i.e. points where the free boundaries have corners, are isolated and finite. If the domain is a strip and there are only two populations, under additional monotonicity assumptions on the boundary data, the free boundary sets {u i > 0}, i = 1, 2, are of class C 1. Lipschitz regularity for free boundary for the obstacle problem associated in dimension 2 (Theorem 8.18): For n = 2, under the additional assumption that p = 1 in (2.5), f i 1 and additional conditions about the regularity of Ω, if (u ε 1,..., uε K ) is a particular solution of (2.4) which satisfies the associated obstacle problem (8.49) with (u 1,..., u K ) the limit as ε 0, then the free boundaries {u i > 0}, i = 1,..., K, are Lipschitz curves of the plane. Free boundary condition (Theorem 9.2): In any dimension, assume that we have 2 populations, H is defined as in (2.5) with ϕ 1, p = 1 and B 1 (x) = B 1 (x) is the Euclidian ball, 0 {u 1 > 0}, e n {u 2 > 0}, and {u 1 > 0} and {u 2 > 0} are of class C 2 in a neighborhood of 0 and e n respectively. Let κ i (0) denote the principal curvatures of {u 1 > 0} at 0, where outward is the positive direction and let κ i (e n ) denote the principal curvatures of {u 2 > 0} at e n where now inward is the positive direction. Then, we have the following relation on the exterior normal derivatives of u 1 and u 2 : u 1 ν(0) u 2 ν(e n ) = n 1 i=1 κ i (0) 0 κ i (0) κ i (e n ) if κ i (0) 0 for some i = 1,..., n 1, and u 1 ν(0) = u 2 ν(e n ) if κ i (0) = 0 for any i = 1,..., n 1. 7

8 4. Existence of solutions This proof follows the same steps as in [30] and it is written below for the reader s convenience. Theorem 4.1. Assume (2.8). Then there exist continuous positive functions u ε 1,..., uε K, depending on the parameter ε, viscosity solutions of the problem (2.4). Proof. The proof uses a fixed point result. Let B be the Banach space of bounded continuous vector-valued functions defined on the domain Ω with the norm ( (u 1, u 2,..., u K ) B := max i sup x Ω For i = 1,..., K, let φ i be the solutions of { φ i = 0 in Ω, (4.1) φ i = f i on Ω. ) u i (x). Let Θ be the subset of bounded continuous functions in Ω, that satisfy prescribed boundary data, and are bounded from above and from below as stated below: Θ = {(u 1, u 2,..., u K ) u i : Ω R is continuous, 0 u i φ i in Ω, u i = f i on ( Ω) 1 }. Notice that Θ is a closed and convex subset of B. Let T ε be the operator that is defined on Θ in the following way: T ε( (u 1, u 2,..., u K ) ) := (v ε 1, vε 2,..., vε K ) if for any i = 1,..., K, vε i is solution to the following problem: (vi ε )(x) = 1 ε (4.2) 2 vε i (x) H(u j )(x) in Ω j i vi ε = f i on ( Ω) 1, where u j, j i are given. Observe that if T ε has a fixed point T ε( (u ε 1, u ε 2,..., u ε K) ) = (u ε 1, u ε 2,..., u ε K) then (u ε 1, uε 2,..., uε K ) is a solution of problem (2.4). In order for T ε to have a fixed point, we need to prove that it satisfies the hypothesis of the Schauder fixed point Theorem, see [23]: (1) T ε (Θ) Θ : 8

9 Classical existence results guarantee the existence of a viscosity solution (v1 ε, vε 2,..., vε K ) of problem (4.2) which is smooth in Ω. Since f i 0 and f i 0, the strong maximum principle implies vi ε > 0 in Ω. This implies that (4.3) v ε i 0 in Ω, and, again from the comparison principle, we have v ε i φ i in Ω. We have proved that T ε( (u 1, u 2,..., u K ) ) Θ. (2) T ε is continuous : Let us assume that ((u 1 ) m,..., (u K ) m ) (u 1,..., u K ) in B meaning that when m tends to +, max (u i) m u i L 0. 1 i K We need to prove that for each fixed ε > 0 T ε( (u 1 ) m,..., (u K ) m ) T ε (u 1,..., u K ) B 0 when m +. Let T ε( (u 1 ) m,..., (u K ) m ) = ( (v ε 1 ) m,..., (v ε K) m ), then if we prove that there exists a constant C ε independent of m, so that we have the estimate, for i = 1,..., K (vi ε ) m vi ε L C ε max (u j ) m u j L, j the result follows. For all x Ω and for fixed i, let ω m be the function ω m (x) = (v ε i ) m (x) v ε i (x), 9

10 and suppose for instance that there exists y Ω such that (4.4) ω m (y) > r 2 D max (u j ) m u j L, j for some large D > 0, where r is such that Ω B r, and B r is the ball centered at 0 of radius r in the Euclidean norm. We want to prove that this is impossible if D is sufficiently large. Let h m be the concave radially symmetric function h m (x) = γ m ( r 2 x 2), with γ m = D max j (u j ) m u j L. Observe that: (a) h m (x) = 0 on B r ; (b) h m (x) r 2 D max j (u j ) m u j L for all x in B r ; (c) 0 = ω m (x) h m (x) on Ω, since (v ε i ) m and v ε i are solutions with the same boundary data. Since we are assuming (4.4), there exists a negative minimum of h m ω m in Ω. Let x 0 Ω be a point where the minimum value of h m ω m is attained. Then h m (x 0 ) ω m (x 0 ) < 0 and (h m ω m )(x 0 ) 0. Then, we have ω m (x 0 ) = ( (v ε i ) m ) (x0 ) v ε i (x 0 ) = 1 ε 2 ( ((v ε i ) m (x 0 ) v ε i (x 0 )) j i H((u j ) m )(x 0 ) v ε i (x 0 ) j i ) (H(u j )(x 0 ) H((u j ) m )(x 0 )) 1 ε 2 ( ((v ε i ) m (x 0 ) v ε i (x 0 )) j i H((u j ) m )(x 0 ) ) vi ε (x 0 )(K 1)C max (u j ) m u j j L (Ω) 10

11 adding and subtracting 1 ε 2 v ε i (x 0) j i H((u j) m )(x 0 ), where C depends on the f j s and ϕ. Then 0 (h m ω m )(x 0 ) 2γ m n 1 ε 2 ( ((v ε i ) m v ε i )(x 0 ) j i H((u j ) m )(x 0 ) ) vi ε (x 0 )(K 1)C max (u j ) m u j j L 2nD max j (u j ) m u j L + 1 ε 2 vε i (x 0 )(K 1)C max (u j ) m u j j L 2nD max (u j ) m u j L + C j ε 2 max (u j ) m u j j L because 0 < h m (x 0 ) < ω m (x 0 ) = ( (v ε i ) m v ε i ) (x0 ) and j i H((u j) m )(x 0 ) 0 and so Taking D = D ε > 1 ( (v ε ε 2 i ) m vi ε ) (x0 ) H((u j ) m )(x 0 ) 0. C 2nε 2, we obtain that j i 0 (h m ω m )(x 0 ) < 0 which is a contradiction. (3) T ε (Θ) is precompact : Let ( (u 1 ) m,..., (u K ) m ) be a bounded sequence in B and let ( (v ε 1 ) m,..., (v ε K) m ) = T ε ( (u 1 ) m,..., (u K ) m ). Then by standard Hölder estimates for viscosity solutions, ( (v ε 1 ) m,..., (v ε K ) m) is bounded in the space of Hölder continuous functions on Ω. Since the subset of Θ of Hölder continuous functions on Ω is precompact in Θ, we can extract from ( (v ε 1 ) m,..., (v ε K ) m) a subsequence which is converging in B. We have proven the existence of a solution (u ε 1,..., uε K ) of (2.4). The same argument as in (1) shows that u ε i > 0 in Ω. This concludes the proof of the theorem. 11

12 5. Uniform in ε Lipschitz estimates In this section we will prove uniform in ε Lipschitz estimates that will imply the convergence, up to subsequences, of the solution (u ε 1,..., uε K ) of (2.4) to a limit function (u 1,..., u K ) as ε 0. We will show that the functions u i s are locally Lipschitz continuous in Ω and harmonic inside their support. Moreover, u i 0 in the ρ-strip of size 1 of the support of u j for any j i, i.e., the supports of the limit functions are at distance at least 1, one from each other. We start by proving general properties of subsolutions of uniform elliptic equations. Lemma 5.1. Let: a) ω be a subharmonic function in B 1, such that a 1 ) ω 1 in B 1 ; a 2 ) ω(0) = m > 0. b) D 0 be a smooth convex set with bounded curvatures κ i ( D 0 ) C 0, i = 1,..., n 1 (like B 1 above). Then, there exists a universal τ 0 = τ 0 (C 0, n, ρ) such that, if the distance d ρ (D 0, 0) τ 0 m, then sup D 0 B 1 ω m 2. Proof. Assume w.l.o.g. that 0 / D 0 and let h be harmonic in B 1 \ D 0 and such that h = 1 on ( B 1 ) \ D 0 h = m 2 on ( D 0 ) B 1. By assumption (b), the set B 1 \ D 0 satisfies an exterior uniform ball condition at any point of D 0 B 1, therefore, by a standard barrier argument, h grows no more than linearly away from D 0 in B 1, i.e., there exist k 1, k 2 > 0 depending on C 0 and n such that, if x B 1 \ D and d(x, D 0 ) k 2, then h(x) k 1 d(x, D 0 ) + m 2. To prove that h(0) < m observe that if τ 0 k 2 c 1, where c 1 is given by (2.2), then d(0, D 0 ) τ 0 m/c 1 k 2 m k 2 and therefore, if in 12

13 addition τ 0 is so small that k 1 c 1 τ 0 1 2, we have h(0) k 1 d(0, D 0 ) + m 2 k 1 c 1 d ρ (0, D 0 ) + m 2 k 1 c 1 τ 0 m + m 2 < m. Hence, we must have sup ( D0 B 1 ) ω m 2, otherwise the comparison principle would imply ω(x) h(x) in B 1 \ D 0, which is a contradiction at x = 0. Lemma 5.2. Let ω be a positive subsolution of a uniformly elliptic equation, (λ 2 I a ij Λ 2 I) a ij D ij ω θ 2 ω in B r. Then there exist c, C > 0 such that Proof. The function ω(0) sup B r ω Ce cθr. g(x) = n i=1 ( ) θ cosh Λ x i is a supersolution of the equation a ij D ij u = θ 2 u. Moreover, using the convexity of the exponential function, it is easy to check that it satisfies g(x) C 1 e cθr for any x B r. Then, the comparison principle implies The result follows taking x = 0. ω(x) sup ω B r g(x) C 1 e cθr for any x B r. The next lemma says that if u ε i attains a positive value σ at some interior point, then all the other functions u ε j, j i, go to zero exponentially in a ρ-ball of radius 1+cσ around that point. Lemma 5.3. Assume (2.8). Let (u ε 1,..., uε K ) be a viscosity solution of the problem (2.4). For i = 1,..., K, σ > 0, and 0 < r < 1 let Γ σ,r i := {y Ω : d ρ (y, supp f i ) 2r, u ε i = σ} 13

14 and m := σ sup Ω f i. Then, there exists a universal constant 0 < τ < 1 such that, in the sets Σ σ,r i,j := { x Ω : d ρ (x, Γ σ,r i ) 1 + τmr 2, d ρ(x, supp f j ) τmr } 4 we have u ε j Ce cσα r β ε, for j i, for some positive α and β depending on the structure of H (p and q). Proof. Let 0 < τ < 1 to be determined. For 0 < r < 1, let us consider the set Σ σ,r i,j above and let x Σ σ,r i,j. We want to show that for j i, we have (5.1) u ε j Cσα r β ε 2 u ε j in B τmr 4 (x) for some α, β > 0. Let us prove it for x such that d ρ (x, Γ σ,r i defined ) = 1 + τmr 2, which is the hardest case. First of all, remark that since d ρ (x, supp f j ) τmr 4, the ball B τmr (x) does not intersect supp f j. Therefore, u ε j (which is eventually zero in B τmr 4 (x) Ωc ) satisfies (5.2) u ε j 1 ε 2 uε j H(u ε k ) Next, the ball B 1 τmr 2 k j in B τmr 4 (x). (x) is at distance τmr from a point y Γσ,r i. Remark that since B 2r (y) supp f i =, the function u ε i (which is eventually equal to zero in B 2r(y) Ω c ) satisfies u ε i 0 4 in B 2r (y). Moreover, since u ε i is subharmonic in Ω, it attains its maximum at the boundary of Ω, so that u ε i / sup Ω f i 1 in Ω. In particular m = σ sup Ω f i (5.3) v(x) := uε i (y + rx), sup Ω f i 1. Set then v 1 and v(0) = u ε i (y)/ sup Ω f i = σ/ sup Ω f i = m and v 0 in B 1. Let ( ) x y D 0 := B 1 r τm, 2 r then the principal curvatures of D 0 satisfy κ i ( D 0 ) C ρ 1 r τm 2 = 2rC ρ 2 rτm < 2rC ρ < 2C ρ. 14

15 Moreover D 0 is at distance τm from 0. Hence, from Lemma 5.1 applied to the function v given by (5.3) with D 0 defined as above, if τ = min{1, τ 0 }, where τ 0 is the universal constant given by the lemma, then there is a point z in B 1 τmr (x) B r(y), such that u ε 2 i (z) σ/2. Next, remark that if x B τmr (x) then 4 B 1 (x) B τmr 4 (z) (since d ρ (x, z) d ρ (x, x) + d ρ (x, z) τmr τmr 2 = 1 τmr 4 ). Let us first consider the case H defined as in (2.6). Then for any x B τmr (x) we have 4 H(u ε i )(x) = sup u ε i u ε i (z) σ B 1 (x) 2, which, with together (5.2), implies (5.1) with α = 1 and β = 0. Next, let us turn to the case H defined as in (2.5). Remark that since z B r (y) and d ρ (y, supp f i ) 2r, we have that B r (z) supp f i = and therefore the function u ε i (which is eventually equal to zero in B r (z) Ω c ) satisfies u ε i 0 in B r(z). This implies that (u ε i )p, p 1, is subharmonic in B r (z) and by the mean value inequality (5.4) ( σ ) p (u ε i ) p dx B s(z) 2 in any Euclidian ball B s (z) B r (z), for any p 1. Since d ρ and the Euclidian distance are equivalent, there is an s τmr such that (5.5) B s (z) B τmr 8 (z) B τmr 4 (z) B 1(x). Moreover, if y B s (z) and x B τmr (x), then 4 that is ρ(y x) ρ(y z) + ρ(z x) + ρ(x x) τmr 8 (5.6) 1 ρ(y x) τmr ( + 1 τmr 2 ) + τmr 4 = 1 τmr 8,

16 Hence, using (5.5), (2.7), (5.6) and (5.4), for all x B τmr (x) we get 4 H(u ε i )(x) = (u ε i ) p (y)ϕ(ρ(y x))dy B 1 (x) B s(z) B s(z) Cσ α r β (u ε i ) p (y)c(1 ρ(y x)) q dy (u ε i ) p (y)c ( τmr ) q dy 8 where α and β depend on p, q and on the dimension n. This and (5.2) imply (5.1). Now, by Lemma 5.2 we get u ε j(x) Ce cσα r β ε for α = α and β = β 2 + 1, and the lemma is proven. Corollary 5.4. Assume (2.8). Let (u ε 1,..., uε K ) be a viscosity solution of the problem (2.4). Let y be a point in Ω such that where m = u ε i (y) = σ, d ρ (y, supp f j ) 1 + τmr, i j and d ρ (y, Ω) 2r, σ sup Ω f i, 0 < r < 1, ε σ 2α r 2β and τ, α and β are given by Lemma 5.3. Then there exists a constant C 0 > 0 such that in B τmr (y) we have 4 (5.7) u ε i C 0 r and (5.8) u ε i 0 as ε 0 uniformly. Proof. First of all, remark that m 1, as u i attains its maximum at the boundary of Ω. Since in addition τ < 1, we have that B τmr 2 (y) B 2r(y) Ω. Therefore, we use (2.4) to estimate u ε i (z), for z B τmr (y). In order to do that, we need to estimate H(uε 2 j )(z) for j i. But H(u ε j )(z) involves points x at ρ-distance 1 from z. Let x be such that d ρ(x, z) 1, then 16

17 d ρ (x, y) 1 + τmr 2. Moreover, since d ρ(y, supp f j ) 1 + τmr, we have d ρ (x, supp f j ) τmr 2. Hence, by Lemma 5.3, for any j i u ε j(x) Ce cσα r β ε for x B 1 (z). From the previous estimate and (2.4), it follows that for z B τmr (y) we have 2 (5.9) 0 u ε i (z) u ε i (z) cσ α r β Ce ε ε 2 u ε i (z) Ce cε 1 2 ε 2 = o(1) as ε 0, for ε σ 2α r 2β. If we normalize the ball B τmr 2 then we have and where u ε i (z) := 2 uε i 0 u ε i (z) τmr 2 uε i (z) j i (y) in a Lipschitz fashion: ( τmr 2 z + y), τmr u ε i (0) = 2 uε i (y) τmr = 2 sup Ω f i, τr 1 ( τmr ) ε 2 H(uε j) 2 z + y τmr 2 uε i (z) 1 1 ( τmr ) ε 2 H(uε j) 2 z + y Ce cε 2 ε 2. j i Then, by the Harnack inequality (see e.g. Theorem 4.3 in [3]), we get sup B 12 (0) u ε i C n inf B 12 (0) uε i + Ce cε 1 2 ε 2 C n 2 sup Ω f i τr for z B 1 (0), + Ce cε 1 2 ε 2 Lipschitz estimates then imply that u ε i C/r in B 1 2 (0) and(5.7) follows. C r. Further, (5.9) implies (5.8). The next lemma says that in a ρ-strip of size 1 of support of the f j s, the function u ε i, i j, decays to 0 exponentially. Lemma 5.5. Assume (2.8). Let (u ε 1,..., uε K ) be a viscosity solution of the problem (2.4). For j = 1,..., K, σ > 0, let Γ σ j := {f j σ} Ω c. Then on the sets {x Ω : d ρ (x, Γ σ j ) 1 r}, 0 < r < 1 17

18 we have u ε i Ce cσα r β ε, for i j, for some positive α and β depending on the structure of H (p and q) and the modulus of continuity of f j. Proof. Let x Ω and y Γ σ j be such that d ρ (x, y) 1 r. We want to estimate H(u ε j )(x), for any x B r (x). Let x B r (x), then 2 2 (5.10) d ρ (x, y) 1 r 2. Let us first consider the case H defined as in (2.6). We have H(u ε j)(x) = sup u ε j f j (y) σ. B 1 (x) Next, let us turn to the case H defined as in (2.5). Let r 0 := min{σ γ, r/4}, for some γ depending on the modulus of continuity of f j, then f j σ/2 in the set B r0 (y) supp f j. Moreover, remark that from (5.10) and r 0 r/4, we have B r0 (y) supp f j B r 4 (y) B r 2 (y) B 1(x), and for any z B r0 (y) supp f j ρ(x z) ρ(x y) + ρ(y z) 1 r 2 + r 0 1 r 4. Therefore, using in addition (2.7), and that, by (2.8), B r0 (y) supp f j c B r0 (y), we get H(u ε j)(x) = (u ε j) p (z)ϕ(ρ(x z))dz B 1 (x) Cσ p r β 0, where β depends on q and on the dimension n. (u ε j) p (z)(1 ρ(x z)) q dz B r0 (y) supp f j ( r ) q (f j ) p (z)c dz B r0 (y) supp f j 4 18

19 Then, for H defined as in (2.5) or (2.6), the function u ε i, i j (which is eventually zero in B r 2 (x) Ωc ) is subsolution of u ε i u ε i Cσ p r β 0 ε 2 in B r (x), where p = 1 and β = 0 in the case (2.6). The conclusion follows as in Lemma The following corollary is a consequence of Lemma 5.3, Corollary 5.4 and Lemma 5.5. Corollary 5.6. Assume (2.8). Let (u ε 1,..., uε K ) be a viscosity solution of the problem (2.4). Then, there exists a subsequence (u ε 1 1,..., uε l K ) and continuous functions (u 1,..., u K ) such that, (u ε l 1,..., uε l K ) (u 1,..., u K ) as l +, a.e. in Ω and the convergence of u ε l i to u i is locally uniform in the set {x Ω : d ρ (x, supp f j ) > 1, j i}. Moreover, we have: i) the u i s are locally Lipschitz continuous in Ω and have disjoint supports, in particular u i 0 in the set {x Ω d ρ (x, supp u j ) 1} for any j i. ii) u i = 0 when u i > 0. Proof. Fix an index i = 1,..., K. Let us denote and Claim 1: u ε i (x) 0 as ε 0 for any x B i. that Ω i := {x Ω d ρ (x, supp f j ) > 1 for any j i}, B i := Ω \ Ω i. Indeed, let x 0 belong to B i, then there exists j i such that d ρ (x 0, supp f j ) < 1. Remark {x Ω d ρ (x, supp f j ) < 1} r,σ>0 {x Ω d ρ (x, Γ σ j ) 1 r}, where Γ σ j = {f j σ}. Therefore, there exist r, σ > 0 such that x 0 {x Ω d ρ (x, Γ σ j ) 1 r}, and by Lemma 5.5 we have that u ε i (x 0) Ce cσα r β ε, for some α, β > 0. Claim 1 follows. 19

20 Claim 2: there exists a subsequence (u ε l i ) l locally uniformly convergent in Ω i as l + to a locally Lipschitz continuous function u i. Fix, 0 < r < 1, and define Ω r i := {x Ω i d ρ (x, Ω) > 2r, d ρ (x, supp f j ) 1 + τr for any j i}, Fix θ < 1 2α and set σ ε = ε θ > 0 and consider τ, α and β as given by Lemma 5.3. Since ε = σε 2α σ 1 θ 2α ε = σε 2α ε θ( 1 θ 2α) and 1 θ 2α > 0, we can fix ε 0 = ε 0 (r) so small that for any ε < ε 0 we have that ε σε 2α r 2β. Then, by Corollary 5.4, the functions v ε i := (u ε i σ ε ) + = (u ε i ε θ ) + are Lipschitz continuous in Ω r i. Indeed, when uε i (x) < εθ, then vi ε (x) = 0. Next, let x such that u ε i (x) > εθ, then vi ε(x) = uε i (x). Set σ = uε i (x), then at those points x, we have that d ρ (x, supp f j ) 1 + τr 1 + mτr, where m = σ/ sup Ω f i 1. Moreover, d ρ (x, Ω) > 2r and ε σε 2α r 2β σ 2α r 2β. We can therefore apply Corollary 5.4 and we get that u ε i (x) C 0 r. This concludes the proof that the functions v ε i are Lipschitz continuous in Ω r i. Therefore, we can extract a subsequence (v ε l i ) l uniformly convergent to a Lipschitz continuous function u i in Ω r i as l +. By the definition of the v i s, this implies that there exists a subsequence (u ε l i ) l uniformly convergent to the same function u i in Ω r i as l +. Taking r 0 and using a diagonalization argument, we can find a subsequence of (u ε i ) ε converging locally uniformly to a Lipschitz function u i in Ω i. This ends the proof of Claim 2. Claims 1 and 2 yield the convergence, up to a subsequence, of u ε i to a continuous function u i which is locally Lipchitz in both Ω i and B i. The fact that the supports of the limit functions are at distance greater or equal than 1, is a consequence of Claims 1 and 2 and Lemma 5.3. Moreover, from the proof of Claim 2 and Corollary 5.4, we infer that the limit function u i is harmonic inside its support, i.e. (ii). To conlude the proof of (i), we just need to prove that u i is Lipschitz in a neighborhood of points belonging to B i = Ω i Ω. Let x 0 Ω i Ω, then 20

21 from Claim 1, u i (x 0 ) = 0. If x 0 {u i > 0}, then in a neighborhood of x 0, u i 0 and of course it is Lipschitz there. On the other hand, if x 0 {u i > 0}, then, since there exists an exterior ρ-tangent ball of radius 1 at any point of Ω i Ω and u i is harmonic inside its support, a barrier argument implies that there exist r 0, C > 0 such that 0 u i (x) = u i (x) u i (x 0 ) C x x 0 for any x B r0 (x 0 ). This concludes the proof of (i). This concludes the proof of the corollary. 6. A semiconvexity property of the free boundaries Let (u 1,..., u K ) be the limit of a convergent subsequence of (u ε 1,..., uε K ), whose existence is guaranteed by Corollary 5.6. For i = 1,..., K, let us denote (6.1) S(u i ) := {x Ω : u i > 0}. (In the next sections, for simplicity this set will be represented by S i.) Then the sets S(u i ) have the following semiconvexity property: Lemma 6.1. Given S(u i ) consider T (u i ) = { x Ω : d ρ (x, S(u i )) 1 } and S (u i ) = { x Ω : d ρ (x, T (u i )) > 1 } Then S(u i ) S (u i ). Proof. We have that S (u i ) S(u i ). To prove the desired inclusion, for σ > 0 consider the sets S σ (u i ) := {x Ω : u i > σ}, T σ (u i ) := {x Ω : d ρ (x, S σ (u i )) 1} and Sσ(u i ) := {x Ω : d ρ (x, T σ (u i )) > 1}. 21

22 Notice that, the union of ρ-balls centered at points in S σ (u i ) coincides with the union of ρ-balls centered at points in S σ(u i ), i.e. a) (T σ (u i )) c = B 1 (x) for x S σ (u i ) and b) (T σ (u i )) c = B 1 (x) for x S σ(u i ). If x S σ (u i ), from (i) of Corollary 5.6 we have that d ρ (x, suppf j ) > 1 for j i, and the locally uniform convergence of u ε i to u i and Lemma 5.3 imply that, up to subsequences, u ε j Ce cσα r β ε in B 1 (x), where 2r = min{d ρ (x, suppf i ), C(d ρ (x, suppf j ) 1)}. Now, the set where u ε j decays is the same if we had considered x S σ(u i ), since from (a) and (b) we have x Sσ(u i )B 1 (x) = x S σ (u i )B 1 (x). Therefore H(uε j ) ε 2 goes to zero as ε goes to zero in S σ(u i ). It follows that u i 0 in S σ(u i ), if S σ(u i ) is not empty. Now, if A is a connected component of S σ (u i ), then there exists a connected component A of Sσ(u i ) such that A A. Since u i is harmonic and non-negative in A, the strong maximum principle implies that u i > 0 in all A, that is A A. We conclude that A = A. Therefore, any connected component of S σ (u i ) is equal to a connected component of Sσ(u i ). Passing to the limit as σ 0, we obtain that any connected component of S(u i ) is equal to a connected component of S (u i ). In particular, S(u i ) S (u i ). From Lemma 6.1 we can conclude that the sets S(u i ) have a tangent ρ-ball of radius 1 from outside at any point of the boundary, as stated in the following corollary. Corollary 6.2. If x 0 S(u i ) Ω there is an exterior tangent ball, B 1 (y) at x 0, in the sense that for x B 1 (y) B 1 (x 0 ), all u j (x) 0 (including u i ). The following two lemmas about the distance function may be known in the literature and we provide the proof here for the reader s convenience. 22

23 Lemma 6.3. Let S be a closed set. Let d ρ (, S) denote the ρ-distance function from S. Then, in the set {x : d ρ (x, S) > 0}, d ρ (, S) satisfies in the viscosity sense d ρ (, S) C d ρ (, S), where C is a constant depending on n, Dd ρ (, S) L and the constant A given in (2.1). Proof. We first prove that there exists a smooth tangent function from above at any point of the graph of d ρ (, S) in the set {d ρ (, S) > 0}. For simplicity we will write d S ( ) instead of d ρ (, S). Let y 0 be a point in the complementary of S. Let x S be a point where y 0 realizes the distance from S. Assume, without loss of generality, that x = 0. Then d ρ (y 0, 0) = ρ(y 0 ) = d S (y 0 ). In particular, the ball B ρ(y0 )(y 0 ) is contained in S c and tangent to S at 0. For any y B ρ(y0 )(y 0 ), we have that d S (y) d ρ (y, 0) = ρ(y), therefore the cone, graph of the function y ρ(y), is tangent from above to the graph of d S ( ) at (y 0, d S (y 0 )). Next, let ψ be a test function touching from below d S ( ) at y 0, then ψ touches from below the function ρ(y) at y 0. In particular, ψ(y 0 ) ρ(y 0 ). Let us compute ρ. Using (2.1), we get which gives D 2 (ρ) = 1 ( ) 1 Dρ Dρ ρ D2 2 ρ2 1 ρ ρ (AI n Dρ Dρ), In particular, This concludes the proof. ψ(y 0 ) ρ C ρ. C ρ(y 0 ) = C d S (y 0 ). Lemma 6.4. Let S be a closed and bounded set. Let us denote by d ρ (, S) the ρ-distance function from S and by (S) 1 the set at ρ-distance 1 from S. Then (S) 1 has finite perimeter. Proof. For simplicity we will write d S ( ) instead of d ρ (, S), as in the previous lemma and first we present an heuristic proof integrating d 2 S over the set {0 < d S < 1}. Since Dd S is bounded, 23

24 from Lemma 6.3, we see that Therefore, integrating d 2 S, we get C d 2 Sdx = = {0<d S <1} {d S =1} d 2 S = 2 Dd S 2 + 2d S d S C. {d S =0} 2Dd S n dh n 1 c 2d S Dd S n dh n 1 + {d S =1} {d S =1} dh n 1 = ch n 1 ({d S = 1}), 2d S Dd S n dh n 1 where n = Dd S / Dd S is the unit exterior normal. This provides un upper bound for H n 1 ({d S = 1}) and concludes the heuristic proof. To make the argument precise, we need to correct the regularity problem over the boundary. For that, consider a smooth function η with compact support in (0, 1) such that 0 η(ξ) ξ for any ξ [0, 1], η(ξ) = ξ for ξ [δ, 1 δ], η c on (0, 1 δ) and η (ξ) c/δ for ξ (1 δ, 1), where δ > 0 is a small parameter. Then, in a weak sense we have (6.2) div(η(d S )Dd S ) = η (d S ) Dd S 2 + η(d S ) d S. Moreover, from Lemma 6.3, in the set {0 < d S < 1} we have η(d S ) d S η(d S ) C d S C in the viscosity sense and therefore in the distributional sense (see, e.g., [24] for the equivalence between viscosity solutions and weak solutions). Therefore, since η(d S ) is a function with compact support in {0 < d S < 1}, we get 0 = div(η(d S )Dd S ) (6.3) = c δ {0<d S <1} {0<d S <1 δ} {1 δ<d S <1} {1 δ<d S <1} η (d S ) Dd S 2 dx + η (d S ) Dd S 2 dx + C Dd S 2 dx + C. {0<d S <1} {1 δ<d S <1} η (d S ) Dd S 2 dx + C η (d S ) Dd S 2 dx + C Now, using the coarea formula and the inequalities above, we get 1 1 H n 1 ({d S = t})dt = 1 Dd S 2 dx C. δ 1 δ δ {1 δ<d S <1} 24

25 Finally, taking the limit as δ 0 + and using the lower semicontinuity of the perimeter with respect to the convergence in measure, we infer that 1 Per({d S = 1}) lim inf δ 0 + δ This concludes the proof of the lemma. 1 1 δ H n 1 ({d S = t})dt C. Corollary 6.5. The sets S(u i ), i = 1,..., K have finite perimeter. Proof. The corollary is an immediate consequence of Lemma 6.1 and Lemma A sharp characterization of the interfaces In Section 5 we proved that the supports of the limit functions u i s are at distance at least 1, one from each other (see Corollary 5.6). In this section we will prove that they are exactly at distance 1, as stated in the following theorem. Theorem 7.1. Assume (2.8) with p = 1 in (2.5). Let (u ε 1,..., uε K ) be a viscosity solution of the problem (2.4) and (u 1,..., u K ) the limit as ε 0 of a convergent subsequence. Let x 0 {u i > 0} Ω, then there exists j i such that (7.1) B 1 (x 0 ) {u j > 0}. Proof. It is enough to prove the theorem for a point x 0 for which S(u i ) has a tangent ρ-ball from inside, since such points are dense on S(u i ). Indeed, let x be any point of S(u i ). Let us consider a sequence of points (x k ) contained in S(u i ) and converging to x as k. Let d k be the ρ-distance of x k from S(u i ). Then the ρ-balls B dk (x k ) are contained in S(u i ) and there exist points y k S(u i ) B dk (x k ) where the x k s realize the distance from S(u i ). The sequence (y k ) is a sequence of points of S(u i ) that have a tangent ρ-ball from the inside and converges to x. Next, remark that from (b) in Corollary 5.6, we have that d ρ (x 0, supp f j ) 1 for any j i. If there is a j such that d ρ (x 0, supp f j ) = 1, then (7.1) is obviously true. Therefore, we can assume 25

26 that d ρ (x 0, supp f j ) > 1 for any j i. Then, for small S > 0 we have that B 1+S (x 0 ) supp f j = and from (2.4), we know that We divide the proof in two cases. a) H(u)(x) = u(y)ϕ ( ρ(x y) ) dy and B 1 (x) b) H(u)(x) = sup u(y). y B 1 (x) u ε j 1 ε 2 uε j H(u ε k ) in B 1+S(x 0 ). k j Proof of case a): Let S(u i ) = {x Ω : u i > 0} as in (6.1). Let B S be a small ρ-ball centered at x 0 S(u i ). Then, as a measure, as ε 0, up to subsequence u ε i BS (x 0 ) u i BS (x 0 ) (that has strictly positive mass, since u i is not harmonic in B S (x 0 )). We bound by below Indeed B 1+S (x 0 ) j i u ε jdx by B S (x 0 ) u ε i dx. (7.2) ε 2 u ε i (x)dx = B S (x 0 ) j i = j i j i = j i ε 2 j i B S (x 0 ) B 1 (x) B S (x 0 ) B 1+S (x 0 ) B 2+S (x 0 ) B 1+S (x 0 ) B 1+S (x 0 ) B 1+S (x 0 ) B 1 (y) u ε i (x)ϕ ( ρ(x y) ) u ε j(y)dydx u ε i (x)χ [0,1] ( ρ(x y) ) ϕ ( ρ(x y) ) u ε j (y)dxdy u ε i (x)χ [0,1] ( ρ(x y) ) ϕ ( ρ(x y) ) u ε j (y)dxdy u ε i (x)ϕ ( ρ(x y) ) u ε j(y)dxdy u ε j(y)dy, where χ [0,1] is the indicator function of the set [0, 1]. 26

27 Therefore, for any small positive S, taking the limit in ε we get u j u i > 0 B 1+S (x 0 ) j i B S (x 0 ) which implies that there exists j i such that u j cannot be identical equal to zero in B 1+S (x 0 ). Since S small is arbitrary, the result follows. The case b) is more involved. We may assume x 0 = 0. Let y 0 be such that B µ (y 0 ) S(u i ) and 0 B µ (y 0 ). By Corollary 6.2 we know that there exists a ρ-ball B 1 (y 1 ) such that B 1 (y 1 ) S(u i ) = and 0 B 1 (y 1 ). have Let us first prove two claims. Claim 1: There exists µ < µ and C 1 > 0 such that in the annulus {µ < ρ(x y 0 ) < µ} we u i (x) C 1 d ρ ( x, Bµ (y 0 ) ). Since any ρ-ball B satisfies the uniform interior ball condition, for any point x B µ (y 0 ) there exists an Euclidian ball B R0 (z 0 ) of radius R 0 independent of x contained in B µ (y 0 ) and tangent to B µ (y 0 ) at x. Let m > 0 be the infimum of u i on the set {x B µ (y 0 ) d(x, B µ (y 0 )) R 0 /2}, where d is the Euclidian distance function, and let φ be the solution of { R0 } φ = 0 in 2 < x z 0 < R 0 φ = 0 on B R0 (z 0 ) φ = m on B R 0 (z 0 ) 2 i.e., for n 3, ( R n 2 ) 0 φ(x) = C(n)m x z 0 n 2 1. Since u i is harmonic in B µ (y 0 ) and u i φ on B R0 (z 0 ) B R 0 2 (z 0 ), by comparison principle u i φ in { R 0 2 < x z 0 < R 0 }. In particular, for any x { R 0 2 < x z 0 < R 0 } and belonging to the segment between z 0 and x, using that φ is convex in the radial direction, φ C(n)(n 2)m ν BR0 (z 0 ) = i R 0 27

28 where ν i is the interior normal at B R0 (z 0 ), and (2.2), we get u i (x) C(n)(n 2)m R 0 d(x, B R0 (z 0 )) = C(n, R 0 )md(x, B µ (y 0 )) C 1 d ρ (x, B µ (y 0 )). Therefore, letting x vary in B µ (y 0 ) we get u i (x) C 1 d ρ (x, B µ (y 0 )) for any x B µ (y 0 ) with d(x, B µ (y 0 )) R 0 2. Using (2.2), Claim 1 follows. Next, let e 0 = y 0 /ρ(y 0 ) and fix σ < µ so small that B σ (σe 0 ) {µ < ρ(x y 0 ) < µ} B 1+δ (y 1 ). For r [σ υ, σ + υ] and small υ < σ, let us define u ε i := inf B r(σe 0 ) uε i and u i := inf u i. B r(σe 0 ) Since for r [σ, σ + υ], B r (σe 0 ) (S(u i )) c and u i 0 on (S(u i )) c, we have (7.3) u i = 0 for r [σ, σ + υ]. By Claim 1, we know that in B σ (σe 0 ) we have u i (x) C 1 d ρ (x, B µ (y 0 )) C 1 d ρ (x, B σ (σe 0 )) = C 1 (σ ρ(x σe 0 )). We deduce that for r [σ υ, σ] u i = inf u i inf C 1(σ ρ(x σe 0 )) = C 1 (σ r). B r(σe 0 ) B r(σe 0 ) From the previous inequality and (7.3), we infer that (7.4) u i C 1 (σ r) +, r [σ υ, σ + υ]. Next, for j i, r [σ υ, σ + υ], let us define ū ε j := sup u ε j and ū j := sup u j. B 1+r (σe 0 ) B 1+r (σe 0 ) 28

29 The functions u ε i and ūε j are respectively solutions of r u ε i 1 ε 2 uε i sup u ε i j B 1 (z i r ) j (7.5) r u ε j 1 ε 2 uε j sup u ε i B 1 ( z r) j where r u = u rr + (n 1) u r = 1 r r n 1 ( r r n 1 u and z i r and z j r are respectively the points where the infimum of u ε i on B r(σe 0 ) and the supremum of u ε j on B 1+r(σe 0 ) are attained. Note that in spherical coordinates r ) u = r u + θ u and that if we are on a point where u attains a minimum value in the θ for a fixed r then θ u 0 and the opposite inequality holds if we are on a maximum point. We also remark that therefore y j r := σe 0 + r r + 1 ( zj r σe 0 ) B r (σe 0 ) B 1 ( z j r), (7.6) sup u ε i u ε i (ȳr) j u ε i. B 1 ( z r) j Moreover, since B 1 (z i r) B 1+r (σe 0 ) and u ε j is a subharmonic function, we have (7.7) sup u ε B 1 (z i r ) j sup u ε j B 1+r (σe 0 ) = sup u ε j B 1+r (σe 0 ) = ū ε j. From (7.5), (7.6) and (7.7), we conclude that ( ) (7.8) r u ε i r ū ε j. j i In other words, for any φ Cc (σ υ, σ + υ), φ 0, we have σ+υ ( u ε i r n 1 ( 1 )) σ+υ ( φ dr ū ε σ υ r r rn 1 j r n 1 ( 1 )) φ dr. σ υ r r rn 1 29 j i

30 Passing to the limit as ε 0 along a uniformly converging subsequence, we get σ+υ ( u i r n 1 ( 1 )) σ+υ ( φ dr ū r r rn 1 j r n 1 ( 1 )) φ dr. r r rn 1 σ υ The linear growth of u i away from the free boundary given by (7.3) and (7.4), implies that r u i develops a Dirac mass at r = σ and σ+υ σ υ u i r σ υ j i ( r n 1 ( 1 )) φ dr > 0, r rn 1 for υ small enough. Hence, r ( j i ūj) is a positive measure in (σ υ, σ + υ) and therefore there exists j i such that u j cannot be identically equal to zero in the ball B 1+σ (σe 0 ). Since σ small is arbitrary, the result follows. 8. Classification of singular points and Lipschitz regularity in dimension 2 In this section we study singular points in dimension 2. We will always assume (2.8) with p = 1 in (2.5). From the results of the previous sections we know that the solutions u ε 1,..., uε K of system (2.4), through a subsequence, converge as ε 0 to functions u 1,..., u K which are locally Lipschitz continuous in Ω and harmonic inside their support. For i = 1,..., K, let us denote the interior of the support of u i by S i as in (6.1) and the union of the interior of the supports of all the other functions by (8.1) C i := j i S j. Since the sets S i are disjoint we have C i = j i S j. From Theorem 7.1 we know that S i and C i are at ρ-distance 1, therefore for any point x S i there is a point y C i such that ρ(x y) = 1. We say that x realizes at y the distance from C i. Definition. A point x S i is a singular point if it realizes the distance from C i to at least two points in C i. We say that x S i is a regular point if it is not singular. Geometrically, we can describe regular and singular points as follows. Let x S i be a singular point and y 1, y 2 C i points where x realizes the distance from C i. Then the balls 30

31 x 0 S 1 S 1 u 1 > 0 Figure 2. Asymptotic cone at x 0 B 1 (y 1 ) and B 1 (y 2 ) are tangent to S i at x. Consider the convex cone determined by the two tangent lines to the two tangent ρ-balls B 1 (y 1 ) and B 1 (y 2 ), which does not intersect the two ρ-balls. The intersection of all cones generated by all ρ-balls of radius 1, tangent at x and with center in C i defines a convex asymptotic cone centered at x, see Figure 2. If x S i is a regular point, then there is only one point y C i where x realizes the distance from C i. In this case, the two tangent balls coincide and therefore, by definition the asymptotic cone at x S i is an half-plane. We will show that at regular points S i is the graph of a differentiable function. If θ [0, π] is the opening of the cone at x, we say that S i has an angle θ at x. Regular points correspond to θ = π. When θ = 0 the tangent cone is actually a semi-line and S i has a cusp at x. We will show, later on in this section, that, assuming additional hypothesis on the boundary data and the domain Ω, the case θ = 0 never occurs and therefore the free boundaries are Lipschitz curves of the plane. Lemma 8.1. Let C = {(x 1, x 2 ) : x 2 α x 1 }, α 0, be the asymptotic cone of S i at 0 S i. Then there exist y 1, y 2 C i such that the balls B 1 (y 1 ) and B 1 (y 2 ) are tangent respectively to the lines x 2 = ±αx 1 at 0. Proof. Let y 1, y 2 B 1 (0) be such that x 2 = αx 1 is a tangent line to B 1 (y 1 ) at 0 and x 2 = αx 1 is a tangent line to B 1 (y 2 ) at 0. Suppose by contradiction that y 1, y 2 / C i. Then, any y C i such 31

32 that ρ(y 0) = 1 must lie in the smaller arc in B 1 (0) between y 1 and y 2. Moreover, there exists δ > 0 such that all ρ-balls B 1 (y) have at most as tangent lines at 0 the lines x 2 = ±(α δ)x 1. Then the asymptotic cone at 0 must contain the cone {(x 1, x 2 ) : x 2 (α δ) x 1 }, which is not possible. Lemma 8.2. Assume that S i has an angle θ (0, π] at x 0 S i. Then, there exists a neighborhood U of x 0, a system of coordinates (x 1, x 2 ) and a locally Lipschitz function ψ : ( r, r) R, for some r > 0, such that in the system of coordinates (x 1, x 2 ), we have that x 0 = (0, 0) and S i U = {(x 1, ψ(x 1 ) : x 1 ( r, r)}. If in addition θ = π, then ϕ is differentiable at 0. Proof. Let C be the convex asymptotic cone of S i at x 0. Let us fix a system of coordinates (x 1, x 2 ) such that the x 2 axis coincides with the axis of the cone and is oriented such that the cone is above the x 1 axis. Then we have that x 0 = (0, 0) and C = {(x 1, x 2 ) : x 2 α x 1 } with α = tan( π θ 2 ). To prove that in this system of coordinates, S i is the graph of a function in a small neighborhood of x 0, it suffices to show that there exists a small r > 0 such that, for any t < r, the vertical line {x 1 = t}, intersects S i B r (0) at only one point. Suppose by contradiction that there exists a sequence (t n ) such that t n 0 as n +, and the line {x 1 = t n } intersects S i B r (0) at two distinct points (t n, a n ) and (t n, b n ) with b n > a n. Assume, without loss of generality, that t n > 0 for any n. By Lemma 8.1 there exist y 1, y 2 C i that realize the distance from 0, and such that B 1 (y 1 ) is tangent to the line {(x 1, x 2 ) : x 2 = αx 1 } at 0 and B 1 (y 2 ) is tangent to {(x 1, x 2 ) : x 2 = αx 1 } also at 0. For instance, in ( ) 1 1 the particular case of the Euclidean norm, we would have y 1 = 1 + α 2, α 1 + α 2 and ( ) 1 1 y 2 = 1 + α 2, α 1 + α 2. In general, what we can say is that the x 2 coordinate of y 1 and y 2 is a negative value c. We have that B 1 (y 1 ) B 1 (y 2 ), since θ > 0. Moreover, S i (B 1 (y 1 ) B 1 (y 2 )) =. Then, both points (t n, a n ) and (t n, b n ) must be above B 1 (y 1 ) B 1 (y 2 ) 32

33 for n large enough. Next, let y a n, y b n C i be points where (t n, a n ) and (t n, b n ), respectively, realize the distance from C i. Then the ρ-balls B 1 (y a n) and B 1 (y b n) are exterior tangent balls to S i at (t n, a n ) and (t n, b n ), respectively. Recall that the ρ-distance between the points (t n, a n ) and (t n, b n ) converges to 0 as n +, and so, the point y a n has to belong to the lower half ρ-ball B 1 (t n, a n ) {x 2 < a n } for n large enough. Indeed, if not the tangent ρ-ball B 1 (y a n) would contain (t n, b n ) for n large enough. Similarly, y b n has to belong to the upper half ρ-ball B 1 (t n, b n ) {x 2 > b n } for n large enough. This implies that the tangent ρ-ball B 1 (y b n) will converge to a tangent ball to S i at 0, B 1 (y b ), with y b {x 2 0}. On the other hand, by the definition of the asymptotic cones, all the centers of the tangent balls at 0 must belong to the set B 1 (0) {x 2 c}, where c < 0 is the x 2 coordinate of the points y 1, y 2 defined above. Therefore, we have reached a contradiction. We infer that there exists r > 0 such that S i is the graph of a function ψ : ( r, r) R. Since S i is a closed set, ψ is continuous. Let us prove that ψ is Lipschitz continuous at 0. If C = {x 2 α x 1 } is the tangent cone of S i at x 0 in the system of coordinates (x 1, x 2 ), then for r > 0 small enough we have {x 2 2α x 1 } S i B r (0) { x 2 α } 2 x 1, that is, for x 1 < r, Therefore, ψ is Lipschitz at 0. α 2 x 1 ψ(x 1 ) = ψ(x 1 ) ψ(0) 2α x 1. Next, assume that θ = π. Then, we have that y 1 = y 2, and x 0 is a regular point. Therefore, B 1 (y 1 ) {x 2 < 0} is the unique tangent ball to the graph of ψ at x 0 = (0, 0). Moreover, the tangent cone is the half plane {x 2 0}. Let us show that ψ is differentiable at 0. Assume by contradiction that there exists a sequence of positive points (x n 1 ) ( r, r) such that xn 1 0 as n + and ψ(x n 1 (8.2) lim ) n + x n = β

34 Since there exists a tangent ball from below to the graph of ψ at 0 contained in {x 2 < 0}, we must have β > 0. For any point (x n 1, ψ(xn 1 )) S i there exists a point y n C i such that B 1 (y n ) is tangent to S i at (x n 1, ψ(xn 1 )). Let y 2 C i be the limit of a converging subsequence of (y n ). Then the ρ-ball B 1 (y 2 ) is an exterior tangent ball at S i at 0. Equation (8.2) gives ψ(x n 1 ) β 2 xn 1 for n large enough, i.e., the points (xn 1, ψ(xn 1 )) of the free boundary are above the line {x 2 = β/2 x 1 }. This implies that y 1 y 2, that is the limit ρ-ball B 1 (y 2 ) must be different from B 1 (y 1 ). This is in contradiction with the fact that x 0 is a regular point. Therefore we must have Similarly, one can prove that ψ(x 1 ) lim = 0. x x 1 ψ(x 1 ) lim = 0. x 1 0 x 1 We conclude that ψ is differentiable at 0 and ψ (0) = 0. Lemma 8.3. Assume that there exists an open set U of R 2 such that any point of U S i is regular. Then U S i is a C 1 -curve of the plane. Proof. Let y 0 S i U. By Lemma 8.2, there exists a differentiable function ψ and a small r > 0, such that, in the system of coordinates (x 1, x 2 ) centered at y 0 and with the x 2 axis in the direction of the inner normal of S i at y 0, S i B r (y 0 ) is the graph of ψ. Moreover, in this system of coordinates, ψ(y 0 ) = ψ (y 0 ) = 0. By Corollary 6.2, there exists a tangent ball from below, with uniform radius, at any point of the graph of ψ. This implies that for any x 0 1 < r, there exists a C2 function ϕ x 0 1 tangent from below to the graph of ψ at x 0 1 and such that ϕ x 0 1 C, for some C > 0 independent of x 0 1. Therefore we have, for any x 1 < r, ψ(x 1 ) ϕ x 0 1 (x 1 ) ϕ x 0 1 (x 0 1) + ϕ x (x 0 1)(x 0 1 x 0 1) C x 1 x = ψ(x 0 1) + ψ (x 0 1)(x 1 x 0 1) C x 1 x

35 Now, let us show that ψ is of class C 1. Fix a point x 0 1 and consider a sequence (xl 1 ) converging to x 0 1 as l +. Let p be the limit of a convergent subsequence of (ψ (x l 1 )). Passing to the limit in l the inequality, we get ψ(x 1 ) ψ(x l 1) + ψ (x l 1)(x 1 x l 1) C x 1 x l 1 2, ψ(x 1 ) ψ(x 0 1) + p(x 1 x 0 1) C x 1 x 0 1 2, for any x 1 < r. Since ψ is differentiable at x 0 1, we must have p = ψ (x 0 1 ). Lemma 8.4. Assume that the supports of the boundary data, f i s, on ( Ω) 1 have a finite number of connected components. Then the sets S i s have a finite number of connected components. Proof. Consider all the connected components of S i, S j i, i = 1,..., K and j = 1, 2, 3,.... Remark that for any i and j S j i {x ( Ω) 1 : f i (x) > 0}. Indeed, if not we would have u i = 0 on S j i and u i 0 in S j i. The maximum principle then would imply u i 0 in S j i, which is not possible. Moreover, by continuity, Sj i must contain one connected component of the set {x ( Ω) 1 : f i (x) > 0}. For this reasons we say that the components of S i reach the boundary of Ω. This implies that the connected components of S i are finite Properties of singular points. We start by proving three lemmas that will allow to estimate the growth of the solutions near the singular points. The first lemma claims that positive functions which are superharmonic (subharmonic) in a cone and vanish on its boundary, have at least (at most) linear growth away from the boundary of the cone far from the vertex, with a slope that degenerate in a Hölder fashion approaching the vertex. The power just depends on the opening of the cone. The second and third lemmas generalize these estimates to domains which are sets of points at ρ-distance greater than 1 from a closed bounded set. Then we prove 35

36 that the set of singularities is a set of isolated points and we give a characterization. For the set S i which has finite perimeter, we denote by S i the reduced boundary, that is the set of points whose blow-ups converge to half-planes and the essential boundary, S i, are all points except points of Lebesgue density zero and one. Moreover, H 1 ( S i \ S i ) = 0. For more details see [1, 22]. Lemma 8.5. Let v be a nonnegative Lipschitz function defined on B 1 R n, such that v is locally a Radon measure on B 1 and such that v smooth on S = {v > 0}. Assume that S is a set of finite perimeter. Then, for every smooth φ with compact support contained in B 1 v φ = B 1 S v v φdx φ dh n 1 S ν S where ν S is the measure-theoretic outward unit normal and S is the reduced boundary. Proof. As a distribution and integrating by parts v φ = B 1 S v φdx = S div(v φ) div( vφ) + vφdx. Applying the generalized Gauss-Green theorem (see [7], and also [1, 22] for more details) we obtain the result. Lemma 8.6. Let θ 0 (0, π]. Let C be the cone defined in polar coordinates by C = {(ϱ, θ) ϱ [0, + ), 0 θ θ 0 }. Let u 1 and u 2 be respectively a superharmonic and subharmonic positive function in the interior of C B 2r0, such that u 1 u 2 = 0 on C B 2r0. Then for any r < r 0 /3 there exist R = R(θ 0, r), and constants c, C > 0 depending on respectively (θ 0, u 1, r 0 ) and (θ 0, u 2, r 0 ), but independent of r, such that for any x [r, 3r] [0, R] we have (1) u 1 (x) cr α d(x, C) (2) u 2 (x) Cr α d(x, C) 36

37 where α is given by 1 + α = π θ 0. Proof. Let us introduce the function (8.3) v(ϱ, θ) := ϱ 1+α sin((1 + α)θ). Notice that v is harmonic in the interior of C, since it is the imaginary part of the function z 1+α, where z = x + iy, which is holomorphic in the set C \ (, 0]. Moreover v is positive inside C and vanishes on its boundary. By a barrier argument, u 1 has at least linear growth away from the boundary of C, meaning for ρ [r 0 /2, 3r 0 /2] (far from the vertex and from B 2r0 ) u 1 (x) kd(x, C), for k = c 0 min x C u 1, and for x {x C : r 0 /2 < x < 3r 0 /2, d(x, C) s 0 } where c 0 d(x, C) s 0 and s 0 depend on r 0 and θ 0. Therefore, we can find a constant c > 0 depending on u 1, r 0 and θ 0, such that u 1 cv on C B r0. Since in addition u 1 cv = 0 on C B r0, the comparison principle implies (8.4) u 1 cv in C B r0. Since v is increasing in the radial direction and if we are near C it is also increasing in the θ direction, for r x 3r, with r such that r r 0 { ( )} 3 and d(x, C) R with R = r min 1, tan θ02, u 1 (x) cv(x) Cr α d(x, C) and (a) follows. To prove (b) similarly, we have (8.5) u 2 Cv in C B r0, 37

38 where C depends on (θ 0, u 2, r 0 ) but it is independent of r. In particular, for r x 3r and d(x, C) R 2 u 2 (x) Cv(x) Cr α d(x, C). Lemma 8.7. Let Ω be an open set, C be a closed subset of Ω and S = {x Ω d ρ (x, C) 1}. Let S 1 be a connected component of S. Assume that S 1 = Γ 1 Γ 2, with Γ 1 Γ 2 = {0} and S 1 has an angle θ 0 (0, π] at 0 S 1. Let u 1 be a superharmonic positive function in S 1 B 2r0 (0), with u 1 = 0 on S 1 B 2r0 (0). Then, there exists a sequence (x h ) Γ 1 of regular points convergent to zero, x h 0 as h 0, and there exist balls B Rh (z h ) S 1 tangent to S 1 at x h, where R h c x h, such that u 1 (x) cr α δ h d(x, B R h (z h )) for any x B Rh (z h ) \ B R h (z h ), 4 where α δ is given by 1 + α δ = π θ 0 δ. Proof. Since θ 0 (0, π] for any 0 < δ < θ 0, there exist r δ > 0 and a cone C 1 δ centered at 0 with opening θ 0 δ such that C 1 δ B r δ (0) S 1 B rδ (0). Take a sequence of points t h Cδ 1 B r δ (0) converging to 0 as h 0. Let { ( )} θ0 δ r h := d(t h, 0) and R h := r h min 1, tan. 2 Then, for h small enough, there exist balls B Rh (s h ) C 1 δ B r δ (0) such that t h B Rh (s h ). Consider a system of polar coordinates (ϱ, θ) centered at 0. Moving the balls B Rh (s h ) along the θ direction until it touches Γ 1, we can find a sequence of regular points x h in that region, such that d(x h, 0) cr h and balls B Rh (z h ) S 1 B rδ (0) such that x h B Rh (z h ). Observe that the center of the ball, z h, remains inside the cone Cδ 1, that is, for h and δ small enough, we 38

39 have that z h Cδ 1 and d(z h, Cδ 1) R h 2. Let us introduce the barrier function φ(x) := m ( ) log 4 log Rh, where m = inf x z h u 1. B Rh (z h ) 4 Then φ satisfies φ = 0 in B Rh (z h ) \ B R h (z h ) 4 φ = 0 on B Rh (z h ) φ = m on B R h 4 (z h ). Since u 1 φ on B Rh (z h ) B R h 4 (z h ) the comparison principle then implies u 1 φ in B Rh (z h ) \ B R h 4 (z h ). If ν 1 is the inner normal vector of B Rh (z h ), then for x B Rh (z h ), φ ν 1 (x) = m R h log 4, and the convexity of φ in the radial direction gives, for any x B Rh (z h ) \ B R h 4 (z h ) u 1 (x) m R h log 4 d(x, B R h (z h )). Let us estimate m. Since d(z h, C 1 δ ) R h 2, we have that d(x, C 1 δ ) R h 4 for any x B R h 4 (z h ). As in Lemma 8.6, consider the harmonic function v(x), introduced in (8.3), defined on the cone C 1 δ (α = α δ) and the comparison principle result stated in (8.4). Then m c { min v min v (z h ) 4 B Rh where c 1 = c 1 (u 1, r δ, θ 0 δ). Then, since r h R h ( r h R h 4, θ ) ( 0 δ 3rh, v 8 4, π )} ( ) αδ +1 3rh = c we conclude that for any x B Rh (z h )\B R h 4 (z h ), u 1 (x) cr α δ h d(x, B R h (z h )). This concludes the proof of the lemma. Lemma 8.8. Let Ω be an open set, C be a closed subset of Ω and S = {x Ω d ρ (x, C) 1}. Let S 1 be a connected component of S. Assume that S 1 has an angle θ 0 [0, π] at 0 S 1. Let u 2 be a subharmonic positive function in S 1 B 2r0 (0), with u 2 = 0 on S 1 B 2r0 (0). 39

40 Then, for any 0 < δ < θ 0, there exists r δ > 0 such that for any r < r δ /5 there exist R = R(θ 0, r), and a constant C > 0 depending on (θ 0 + δ, u 2, r δ ), but independent of r, such that (8.6) u 2 (x) Cr β δ d(x, S 1 ) for any x (B 3r (0) \ B r (0)) where β δ is given by 1 + β δ = π θ 0 + δ. { x S 1 : d(x, S 1 ) R } 4 Proof. For any δ > 0, there exist r δ > 0, a cone C 2 δ centered at 0 and with opening θ 0 + δ such that S 1 B rδ (0) C 2 δ B r δ (0). Take any r < r δ and let y S (B 3r (0) \ B r (0)) and r y := d(y, 0) (r, 3r). Since S is at ρ-distance 1 from C, for any point of the boundary of S 1 there exists an exterior tangent ρ-ball of radius 1. This implies that for r small enough, there exists w y such that the Euclidian ball B Ry (w y ) is contained in the complement of S and y B Ry (w y ), where R y is defined by { ( )} θ0 + δ R y = r y min 1, tan. 2 Let us take now as barrier the function ψ(x) := M ( ) wy x log 3 log R 2 y with M = sup B 32 Ry (wy) u 2. Then ψ satisfies ψ = 0 ψ = M in B 3Ry (w y ) \ B Ry (w y ) 2 on B 3Ry (w y ) 2 ψ = 0 on B Ry (w y ). Using the comparison principle with u 2, the concavity of ψ in the radial direction gives that for any x B 3Ry (w y ) \ B Ry (w y ) 2 u 2 M R y log( 3 2 )d(x, B R y (w y )). 40

41 Let us estimate M. Consider again a system of polar coordinates (ϱ, θ) centered at 0 and the harmonic function v(x), introduced in (8.3), defined on the cone C 2 δ (α = β δ). By definition of v, R y, and taking into account (8.5), for δ, r small enough, M C max B 3Ry 2 ( v Cv (w y) 4r y, θ 0 + δ 2 ) = C 1 (4r y ) β δ+1 = C 1 r β δ y R y min{1, tan( θ 0+δ 2 )} we see that for any x B 3Ry (w y ) \ B Ry (w y ) and belonging to the segment y + s(y w y ), 2 s ( 0, 1 2), we have (8.7) u 2 (x) CMd(x, B Ry (w y )) = CMd(x, S 1 ) Cr β δ y d(x, S 1 ). Letting the tangent ball moving along S 1 (B 3ry (0) \ B ry (0)), we get (b). Lemma 8.9. Assume (2.8) with n = 2 and p = 1 in (2.5). Assume in addition that the supports on Ω of the boundary data f i s have a finite number of connected components. Let (u ε 1,..., uε K ) be a viscosity solution of the problem (2.4) and (u 1,..., u K ) the limit as ε 0 of a convergent subsequence. The set of singular points of Ω is a set of isolated points. Proof. Suppose by contradiction that there exists a sequence of distinct singular points (y k ) k N such that y k S j and y k y Ω as k +. Since by Lemma 8.4 the connected components of the sets S i, i = 1,..., K are finite, we may assume without loss of generality that the points y k belong to the same connected component of S j, which we denote by S 1 j. If there exists θ max < π such that S 1 j has an angle smaller than θ max at y k for any k, then, there exists k such that starting from y k, after a finite number of singular points S 1 j would be an isle and not reach the boundary. Therefore we would have u j = 0 on Sj 1 and u j = 0 in Sj 1, and the maximum principle would imply u j 0 in Sj 1, which is a contradiction. We infer that there exists a k N such that the angle at y k is close to π. In particular, if x k 1 and xk 2 are points in C j that realize the ρ-distance from S j at y k, then ρ-distance between x k 1 and xk 2 is less than one. 41

42 Next, suppose that x k i and xk 2 belong to the same connected component of S i, for some i j. Then, by Theorem 7.1 we know that S i B 1 (y k ) has to contain the arc of the unit ρ-ball between x k 1 and xk 2. If not, there would be points in the curve connecting xk 1 and xk 2 which do not realize the distance from C i. Any point inside this arc is a regular point at ρ-distance 1 from y k. Consider any of them, for instance the middle point of the arc, denoted by x k. We want to compare the mass of the Laplacian of u i at x k with the mass of the Laplacian at u j at y k, across the free boundaries. Let us first assume H defined as in (2.5). For σ < 1 8 d ρ(x k 1, xk 2 ) let us define D σ (x k ) := {x B σ (x k ) d(x, C i ) σ 2 }, where C i is the asymptotic cone to S 1 i at x k. Note that, since x k is a regular point, C i is the tangent line to S 1 i at x k, and so C i has opening π. Let (D σ (x k )) 1 be the set of points at ρ-distance less than 1 from D σ (x k ), then we have that (8.8) u i u j D σ(x k ) j i (D σ(x k )) 1 as in (7.2) with (D σ (x k )) 1 in place of B 1+S (x 0 ). By the Hopf Lemma, we obtain (8.9) D σ(x k ) u i = S i D σ(x k ) u i ν i dh c H( S i D σ (x k )) = Cσ where ν i is the inner normal vector. Now we estimate (D σ(x k )) 1 u j. From Corollary 6.5 we know that S j is a set of finite perimeter. Therefore by Lemma 8.5 and Lemma 8.8 we obtain the following estimate (8.10) u j = (D σ(x k )) 1 S 1 j (Dσ(x k)) 1 u j ν S 1 j dh Cσ β δ H( S 1 j (D σ (x k )) 1 ) where ν Sj is the measure-theoretic inward unit normal to S 1 j and β δ > 0. Since, for some constant c S 1 j (D σ (x k )) 1 S 1 j B cσ (y k ), 42

43 by (2.2), there exists c 2, that for simplicity we will still name c, such that Sj 1 (D σ(x k )) 1 Sj 1 B cσ(y k ). Then (8.11) H( Sj 1 (D σ (x k )) 1 ) Per( Sj 1 B cσ (y k )). To estimate Per( Sj 1 B cσ(y k )), consider (6.2) in the distributional sense. Then, take a smooth function 0 φ 1 with compact support contained in B cσ (y k ) {x : 0 < d(x, S i ) < 1} and such that φ 1 on the set B cσ (y k ) {x : 1 δ < d(x, S i ) < 1 ɛ}, for 0 < ε < δ and δ as introduced in the definition of η in the proof of Lemma 6.4. Then, for d Si ( ) = d ρ (, S i ) we have that 0 = B cσ(y k ) {x:0<d Si <1} div(η(d Si )Dd Si )φ = + B cσ(y k ) {x:0<d Si <1} B cσ(y k ) {x:0<d Si <1} B cσ(y k ) {x:0<d Si <1} η (d Si ) Dd Si 2 φdx η(d Si ) d Si φ η (d Si ) Dd Si 2 φdx + Cσ. Proceeding as in Lemma 6.4 we obtain that (8.12) Per( Sj 1 B cσ (y k )) Cσ. Putting together (8.8), (8.9), (8.10), (8.11) and (8.12) we obtain Cσ 1+β δ Cσ and we get a contradiction for σ small enough. In the case (2.6) the proof follows the same steps using (7.8). Therefore we must have that x k 1 and xk 2 belong to different components of C j for any k k. In particular, since the distance between them is less than one, they must belong to two different components of the same population. Suppose that x k 1 S1 i and x k 2 S2 i, for i j. Consider the consecutive two points x k+1 1 and x k+1 2 which realize the distance at y k+1, and again belong to two different components of C j. Since S 1 j (to which y k belongs) and S2 i reach the boundary of Ω, the point x k+1 2 must belong to a connected component different from Si 1. Iterating the 43

44 procedure, we construct a sequence of distinct points belonging to connected components, each different from the others. This is in contradiction with Lemma 8.4. We conclude that singular points are isolated. Theorem Assume (2.8) with n = 2 and p = 1 in (2.5). Let (u ε 1,..., uε K ) be a viscosity solution of the problem (2.4) and (u 1,..., u K ) the limit as ε 0 of a convergent subsequence. For i j, let x 0 S i Ω and y 0 S j Ω be points such that S i has an angle θ i [0, π] at x 0, S j has an angle θ j [0, π] at y 0 and ρ(x 0 y 0 ) = 1. Then we have (8.13) θ i = θ j. If x 0 S i Ω and y 0 S j Ω, then (8.14) θ i θ j. Proof. By Lemma 8.4, the connected components of the sets S i s are finite. Assume x 0 Ω and y 0 Ω. Without loss of generality we can assume that x 0 = 0. It suffices to show the theorem for y 0 belonging to a region that is side by side with S i, in the sense that 0 is the limit as h 0 of interior regular points x h S i Ω with the property that x h realizes the distance from S j at y h S j Ω interior points, with y h y 0 as h 0. Let C i be the asymptotic cone at 0. Let us first suppose for simplicity that S i and S j are locally a cone around 0 and y 0 respectively. In particular, θ i, θ j > 0. We will explain later on how to handle the general case. Proof of Theorem 8.10 when S i and S j are locally cones. We assume that there exists r 0 > 0 such that S i B 2r0 = C i B 2r0, where B 2r0 is the Euclidian ball centered at 0 of radius 2r 0. When x 0 Ω we are just interested in the side of the cone C i contained in Ω. If (ϱ, θ) is a system of polar coordinates in the plane centered at zero, we may assume that C i is the cone given by C i = {(ϱ, θ) ϱ [0, + ), 0 θ θ i }. 44

45 Let us first consider the case (2.6). Let us assume that x h = (2r h, 0), with r h > 0. We know that r h 0 as h 0, then we can fix h so small that r h < r 0 /3. By Lemma 8.6 applied to u 1 = u i, we have (8.15) u i (x) crh α d(x, S i) for any x [r h, 3r h ] [0, R h ], where (8.16) 1 + α = π 1. θ i Now, we repeat an argument similar to the one in the proof of Theorem 7.1. We look at inf u i in small circles of radius r that go across the free boundary of u i and we look at sup u j in circles of radius r + 1 across the free boundary of u j, then we compare the mass of the correspondent Laplacians. Precisely, there exists a small σ > 0 and e S i such that B σ (e) [r h, 3r h ] [0, R h ] and x h B σ (e). In particular, in B σ (e) the function u i satisfies (8.15). For υ < σ and r [σ υ, σ + υ], we define (8.17) u i := inf u i and ū j := sup u j. B r(e) B 1+r (e) In what follows we denote by C and c several constants independent of h. For r [σ υ, σ], by (8.15) we have u i inf B crα h d(x, S i) inf r(e) B Crα h d ρ(x, S i ) Crh α (σ r). r(e) For r [σ, σ + υ], the ball B r (e) goes across S i, therefore we have u i = 0. Hence (8.18) u i (r) Crh α (σ r) for r [σ υ, σ] u i (r) = 0 for r [σ, σ + υ]. Next, let us study the behavior of ū j. First of all, let us show that (8.19) d ρ (e, S j ) = ρ(e y h ) = 1 + σ. Since d ρ (e, S i ) = σ and d ρ (S i, S j ) 1, it is easy to see that d ρ (x, S j ) 1 + σ. The function ρ is also called a Minkowski norm and from known results about Minkowski norms, if we denote by T the Legendre transform T : R n R n defined by T (y) = ρ(y)dρ(y), then T is a bijection with 45

46 inverse T 1 (ξ) = ρ (ξ)dρ (ξ), where ρ is the dual norm defined by ρ (ξ) := sup{y ξ y B 1 }. Now, the ball B 1 (y h ) is tangent to S i at x h and therefore is also tangent to B σ (e) at x h. This implies that Dρ(e x h ) = Dρ(x h e) = Dρ(x h y h ). Consequently we have e x h = T 1 (T (e x h )) = T 1 (σdρ(e x h )) = T 1 (σdρ(x h y h )) We infer that = σt 1 (T (x h y h ) = σ(x h y h ). (8.20) e = x h + σ(x h y h ) and ρ(e y h ) = (1 + σ)ρ(x h y h ) = 1 + σ, which proves (8.19). As a consequence B 1+r (e) S j = for r [σ υ, σ), while if r (σ, σ +υ] then B 1+r (e) S j and B 1+r (e) enters inside S j at ρ-distance at most r σ from the boundary of S j. In particular we have (8.21) ū j = 0 for r [σ υ, σ]. Next, if θ j is the angle of S j at y 0, let β be defined by (8.22) 1 + β = π θ j 1. Remark that y h is at ρ-distance 2r h from y 0. Again by Lemma 8.6 applied to u 2 = u j, (after a rotation and a translation), we have the following estimate u j (x) Cr β h d(x, S j) Cr β h d ρ(x, S j ), in a neighborhood of y h. As a consequence, recalling in addition that the ball B 1+r (e) enters in S j at ρ-distance r σ from the boundary, for r [σ, σ + υ] we get ū j = The last estimate and (8.21) imply sup u j Cr β h (r σ). B 1+r (e) (8.23) ū j (r) Cr β h (r σ)+, for r [σ υ, σ + υ]. 46

47 Now, we want to compare the mass of the Laplacians of u i and ū j. Define as in (8.17) u ε i := inf B uε i, ū ε k := sup u ε k, k i. r(e) B 1+r (e) For σ and υ small enough, the ball B r (e) is contained in Ω for any r σ + υ, and thus u ε i = 1 ε 2 uε i H(u ε k ) k i in B r+σ(e). On the other hand, since x h is an interior regular point that realizes its distance from S j at an interior point, y h, its distance from the support of the boundary data f k is greater than 1, for any k i. We infer that, for σ and υ small enough and r σ + υ, u ε k 1 ε 2 uε k H(u ε l ) l k in B 1+r(e). Hence, arguing as in the proof of Theorem 7.1, we see that (8.24) r u ε i k i r ū ε k in (σ υ, σ + υ), where r u = 1 r ( ) r r u r. Since xh is a regular point of S i that realizes the distance from S j at y h C i, the ball B 1+σ+υ (e) does not intersect the support of the functions u k for k j and small υ and σ. Therefore, multiplying inequality (8.24) by a positive test function φ C c (σ υ, σ + υ), integrating by parts in (σ υ, σ + υ) and passing to the limit as ε 0 along a converging subsequence, the only surviving function on the right-hand side is ū j and we get (8.25) σ+υ σ υ ( u i r ( )) 1 σ+υ ( r r r φ dr ū j r ( )) 1 σ υ r r r φ dr. Let us choose a function φ which is increasing and (σ υ, σ) and decreasing in (σ, σ + υ) and hence with maximum at r = σ, and let us estimates the left and the right hand-side of the last 47

48 inequality. Estimates (8.18) imply that u i r (σ ) Crh α. Therefore, for small υ we have σ+υ ( u i r ( )) 1 σ σ υ r r r φ u dr = i σ υ r r ( ) 1 r r φ dr σ ( ) ui = σ υ r (σ ) + o σ r (1) r ( ) 1 r r φ dr σ ( u i φ σ υ r (σ ) r 1 ) r φ dr σ ( φ o υ (1) σ υ r + 1 ) r φ dr u [ ( )] i σ r (σ ) φ(σ) φ(σ) log σ υ [ ( )] σ o υ (1) φ(σ) + φ(σ) log σ υ (Cr α h o υ(1))φ(σ). Similarly, using (8.23) and integrating by parts, we get σ+υ ( ū j r ( )) 1 r r r φ dr (Cr β h + o υ(1))φ(σ). σ υ From the previous estimates and (8.25), letting υ go to 0, we obtain and therefore, for h small enough r α h Crβ h, β α. Recalling the definitions (8.16) and (8.22) of α and β respectively, we infer that θ i θ j. This proves (8.14). If x 0 = 0 is an interior point of Ω, exchanging the roles of u i and u j, we get the opposite inequality θ j θ i, and this proves (8.13) for H defined as in (2.6). Next, let us turn to the case (2.5). Again we compare the mass of Laplacians of u i and u j across the free boundaries. For σ < r h let us define (8.26) D σ (x h ) := {x B σ (x h ) d(x, S i ) σ 2 }. 48

49 Then, if we denote by (D σ (x h )) 1 the sets of points at ρ-distance less than 1, we have that (8.27) D σ(x h ) u i k i (D σ(x h )) 1 u k, as in (7.2) with (D σ (x h )) 1 in place of B 1+S (x 0 ). By Lemma 8.6 the normal derivative of u i with respect to the inner normal ν i, at any point on the boundary C i with distance to the vertex between r h and 3r h is greater than cr α h, then Remark that D σ(x h ) u i = C i D σ(x h ) u 2rh +Cσ i dh c rh α ν dr = Crα h σ. i 2r h cσ (D σ (x h )) 1 S j B cσ (y h ) S j therefore, for σ small enough, again from Lemma 8.6 we have u j Cr β h σ. (D σ(x h )) 1 Then for r h small enough we obtain that β α and therefore θ i θ j. If x 0 = 0 is an interior point of Ω, exchanging the roles of u i and u j we get the opposite inequality θ j θ i. This concludes the proof of the theorem in the particular case in which S i and S j are locally a cone around 0 and y 0 respectively. We are now going to explain how to adapt the proof in the general case. Proof of Theorem 8.10 in the general case. If θ i = 0, then θ i θ j. Assume θ i (0, π] and θ j [0, π], then for any 0 < δ < θ i, there exist r δ > 0, a cone C i δ centered at 0 and with opening 49

50 θ i δ, and a cone C j δ centered at y 0 and with opening θ j + δ such that Cδ i B r δ (0) S i B rδ (0) and S j B rδ (y 0 ) C j δ B r δ (y 0 ). Let (x h ) h be the sequence of regular points on S i Ω given by Lemma 8.7 (consider Γ 1 the closest side to S j ) and let r h = d(0, x h ). Denote by y h the point on S j Ω at ρ-distance 1 from x h. Then, d ρ (y h, y 0 ) cr h. Now, the proof of the theorem proceeds like in the previous case and we can compare the mass of the laplacians across the free boundaries of u i and u j. Let us first consider the case (2.5). For σ > 0 take D σ (x h ) and (D σ (x h )) 1 as defined as in (8.26). For σ small enough, by Lemma 8.9 S i D σ (x h ) does not contain singular points and by Lemma 8.3 it is a C 1 curve of the plane. By Lemma 8.7 Remark that D σ(x h ) u i = S i D σ(x h ) u i ν i dh Cr α δ h σ. (D σ (x h )) 1 S j B cσ (y h ) S j therefore, for σ small enough, from Lemma 8.8, as in the proof of Lemma 8.9, we have u j Cr β δ h σ. (D σ(x h )) 1 Then for h small enough, we obtain that and therefore β δ α δ θ i θ j. If x 0 = 0 is an interior point of Ω, exchanging the roles of u i and u j we get the opposite inequality θ j θ i. Next, let us turn to the case (2.6). Then, we define, for r [R h υ, R h + υ], u i := inf u i and ū j := sup u j. B r(z h ) B 1+r (z h ) 50

51 Arguing as before, and using the Lemma 8.7 we get β δ α δ, and therefore, letting δ go to 0, we finally obtain θ i θ j. Remark in particular that if θ i > 0 then θ j > 0. If x 0 = 0 is an interior point of Ω, exchanging the roles of u i and u j we get the opposite inequality θ j θ i. An immediate corollary of Theorem 8.10 is the C 1 -regularity of the free boundaries when K = 2 and under the following additional assumptions on Ω, f 1 and f 2 : (8.28) Ω := {(x 1, x 2 ) R 2 g(x 2 ) x 1 h(x 2 ), x 2 [a, b]}, b a 4 where (8.29) { g, h : [a, b] R are Lipschitz functions with m 2 g m 1 M 2 h M 1, M 2 m 1 + 4; the boundary data are such that f 1 1, f 2 0 on {x 1 g(x 2 )}, f 1 0, f 2 1 on {x 1 h(x 2 )}, (8.30) f 1 is monotone decreasing in x 1 on {x 2 a} {x 2 b}, f 2 is monotone increasing in x 1 on {x 2 a} {x 2 b}. These assumptions imply that u 1 and u 2 are monotone increasing in the x 1 direction. Then we have the following Corollary Assume (2.8) with p = 1 in (2.5). Assume in addition K = n = 2, (8.28), (8.29) and (8.30). Then the sets S i, i = 1, 2, are of class C 1. Proof. We know that the sets S i are curves of the plane at ρ-distance 1, one from each other. Suppose by contradiction that S 1 has an angle θ < π at y 0. In particular, there exist two ρ-balls of radius 1, centered at two points z, w S 2 that are tangent to S 1 at y 0. Then, by the monotonicity property of the u i s and Theorem 7.1, the arc of the ρ-ball of radius 1 centered 51

52 at y 0 between the points z and w must be all in S 2. This means that any point inside this arc, which is a regular point of S 2, is at ρ-distance 1 from the singular point y 0 S 1. This contradicts Theorem We have shown that any point of the free boundaries is regular. Then by Lemma 8.3 the free boundaries are of class C 1. This concludes the proof. Another corollary of Theorem 8.10 is that the number of singular points is finite. Corollary Assume (2.8) with n = K = 2 and p = 1 in (2.5). Assume in addition that the supports on Ω of the boundary data f 1 and f 2 have a finite number of connected components. Then singular points form a finite set. Proof. From Lemma 8.4, S 1 and S 2 have a finite number of connected components. Moreover, we recall that any connected component has to reach the boundary. Let x 0 be a singular point belonging to the boundary of the support of one of the limit functions u i. W.l.o.g. let us assume x 0 S 1. Let y 1, y 2 S 2 two different points where x 0 realizes the distance from S 2, (y 1, y 2 B 1 (x 0 ) S 2, see Figure 3). We can choose y 1 such that B 1 (x 0 ) is the limit as k + of balls B 1 (x k ) with x k S 1, tangent to points y k S 2 with y k y 1 and x k x 0 as k +. Theorem 8.10 implies that S 2 has an angle at y 1 and y 2 and the intersection of the arc on B 1 (x 0 ) between y 1 and y 2 with C 1 must have empty interior. This means that near y 1 there are points on S 2 outside B 1 (x 0 ). These points are at distance greater than 1 from x 0 and from any other point of S 1 close to x 0 and must realize the distance from S 1 outside B 1 (y 1 ), see Figure 3. Therefore if we take a sequence z k of such points converging to y 1 and we consider the corresponding tangent balls centered at points that are in S 1 where the z k s realize the distance, we obtain a second tangent ball B 1 (x 1 ) for y 1 with x 1 x 0. Now, let us denote by S1 1 the connected component of S 1 whose boundary contains x 0. Remember that since S 1 and S 2 are at ρ-distance 1, we have u 1 0 in B 1 (y 1 ) B 1 (y 2 ). 52

53 S 2 1 x 1 z k y 1 y 2 x 0 S 2 Forbidden arc S 1 1 S 1 1 Figure 3. Forbidden arc Moreover, since the connected components of S 2 whose boundaries contain y 1 and y 2 must reach the boundary of Ω, they separate the components of S 1 whose boundaries contain x 0 and x 1. Therefore x 1 must belong to the boundary of different components of S 1. The same argument that we have used for x 1 and x 0 proves also that y 1 and y 2 must belong to the boundary of different components of C 1. We conclude that a singular point x 0 of S 1 involves at least four different connected components and there correspond to it another singular point, x 1, belonging to a different component of S 1 (see Figure 4). Assume w.l.o.g. that x 1 S1 2. Since all the connected components must reach the boundary of Ω, x 1 is the only singular point of S 2 1 corresponding to a singular point of S1 1. Since the connected component of S 1 are finite, we infer that there is a finite number of singular points on S1 1. This argument applied to any connected component of S 1 shows that singular points of S 1 are finite. This concludes the proof of the theorem Lipschitz regularity of the free boundaries. In this section, we will show, under some additional assumptions on the domain Ω and the boundary data f i, that we can construct a 53

54 S 2 1 x 1 S 1 2 y 1 y 2 S 2 2 x 0 S 1 1 Figure 4. A singular point involving four components solution of problem (2.4) such that the free boundaries S i of the limiting functions have the following properties: if S i has an angle θ at a singular point, then θ > 0. This result can be rephrased by saying that the free boundaries are Lipschitz curves of the plane. Let us make the assumptions precise. We assume that the domain Ω has the property that for any point of the boundary there are tangent ρ-balls of radius 1 + η, with η > 0 contained in Ω and in its complementary. Precisely: Ω is a bounded domain of R 2 ; (8.31) η > 0 such that x Ω, B 1+η (y), B 1+η (z) such that x B 1+η (y) B 1+η (z), B 1+η (y) Ω, and B 1+η (z) Ω c. On the boundary data f i, i = 1,..., K, we assume, f i 1 in supp f i ; c > 0 s. t. x Ω supp f i, B r (x) supp f i c B r (x), (8.32) d ρ (supp f i, supp f j ) 1, i j, d ρ (supp f i Ω, supp f i+1 Ω) = 1, where f K+1 := f 1 ; Γ i := supp f i Ω is a connected (C 2 -) curve of Ω. We are going to build a solution of (2.4) such that the support of any limiting function u i contains a full neighborhood of Γ i in Ω with Lipschitz boundary. Then we prove that the free 54

55 µ i f i 1 y 0 y i 1 z i 1 µ A i µ z i 2 y i Figure 5. Construction of obstacle boundaries are Lipschitz. In order to do it, we first prove the existence of a solution (u ε 1,..., uε K ) of an obstacle problem associated to system (2.4). Then we show that the functions u ε i s never touch the obstacles, implying that (u ε 1,..., uε K ) is actually solution of (2.4). We consider obstacle functions ψ i, for i = 1,..., K defined as follows. Let y i 1, yi 2 be the endpoints of the curve Γ i. For 0 < µ < λ < 1, we set: Γ µ i := {x Ωc d(x, Γ i ) = µ}, Γ µ,λ i := {x Γ µ i d(x, yi 1), d(x, y i 2) λ}. For µ and λ small enough, Γ µ,λ i is a C 1,1 curve of Ω c with endpoints z i 1, zi 2 such that d(zi l, yi l ) = λ, l = 1, 2. We finally set ( (8.33) A i := {x Ω d(x, Γ µ,λ i ) < λ} = Ω Remark that A i = Γ i ( A i Ω), x Γ µ,λ i ) B λ (x). 55

56 where A i Ω is given by the union of two arcs contained respectively in the balls B λ (z i 1 ) and B λ (z i 2 ), and a curve contained in the set of points of Ω at distance λ µ from Γ i, (see Figure 5). Denote by αl i the angle of A i at yl i, l = 1, 2. Remark that { αl i (8.34) π 2 + o λ(1) if µ 0 αl i 0 if µ λ, where o λ (1) 0 as λ 0. We take as obstacles the functions ψ i : (Ω) 1 R defined as the solutions of the following problem, for i = 1,..., K, (8.35) ψ i = 0 in A i ψ i = f i on ( Ω) 1 ψ i = 0 in Ω \ A i. In this section we deal with the solution (u ε 1,..., uε K ) of the following obstacle system problem: for i = 1,..., K, (8.36) u ε i ψ i in Ω, u ε i (x) 1 ε 2 uε i (x) H(u ε j)(x) in Ω, j i u ε i (x) = 1 ε 2 uε i (x) H(u ε j)(x) in {u ε i > ψ i} j i u ε i = f i on ( Ω) 1. In the whole section we make the following assumptions: ε > 0, (8.31) and (8.32) hold true, (8.37) H is either of the form (2.5) with p = 1, or (2.6) and (2.7) holds true; For i = 1,..., K, A i and ψ i are defined by (8.33) and (8.35) respectively. Theorem Assume (8.37). Then, there exist continuous positive functions u ε 1,..., uε K, depending on the parameter ε, viscosity solutions of the problem (8.36). In particular (8.38) u ε i (x) = 1 ε 2 uε i (x) j i H(u ε j)(x) in Ω \ A i. Moreover, for i = 1,..., K, (8.39) u ε i 0 in Ω, 56

57 in the viscosity sense. Proof. The proof of the existence of a solution (u ε 1,..., uε K ) of (8.36) is a slightly modification of the proof of Theorem 4.1. Here Θ = {(u 1, u 2,..., u K ) u i : Ω R is continuous, ψ i u i φ i in Ω, u i = f i on ( Ω) 1 }. In the set Ω \ A i, we have that u ε i > 0 = ψ i which implies (8.38). Inequality (8.39) is a consequence of the following facts: in the set {u ε i > ψ i} we have u ε i = 1 u ε ε 2 i j i H(uε j ) 0; in the interior of the set {u ε i = ψ i }, u ε i = ψ i = 0; the free boundaries {u ε i > ψ i } have locally finite n 1-Hausdorff measure, see [2]. Theorem Assume (8.37). Let (u ε 1,..., uε K ) be viscosity solution of the problem (8.36). Then, there exists a subsequence (u ε l 1,..., uε l K ) and continuous functions (u 1,..., u K ) defined on Ω, such that (u ε l 1,..., uε l K ) (u 1,..., u K ) as l +, a.e. in Ω and the convergence of u ε l i to u i is locally uniform in the support of u i. Moreover, we have: i) the u i s are locally Lipschitz continuous in Ω, in particular, there exists C 0 > 0 such that, if d ρ (x, Ω) r, then (8.40) u i (x) C 0 r. ii) the u i s have disjoint supports, more precisely: u i 0 in the set {x Ω d ρ (x, supp u j ) 1} for any j i. iii) u i = 0 when u i > 0. iv) u i ψ i in Ω. v) u i = f i on Ω. Proof. The convergence theorem is again a consequence of Lemma 5.3, Corollary 5.4 and Lemma 5.5 which hold true with supp f i and supp f j replaced respectively by supp ψ i = A i and 57

58 supp ψ j = A j (in Lemma 5.3 and Corollary 5.4), and Γ σ j defined as the set {ψ j σ} (in Lemma 5.5). Estimates (5.7) of Corollary 5.4 in particular imply (8.40). Property (iv) is an immediate consequence of u ε i ψ i in Ω. Finally, (v) is implied by the fact that ψ i u ε i φ i in Ω, and φ i = ψ i = f i on Ω, where φ i is given by (4.1). As proven in Corollary 6.2, one can show that the free boundaries satisfy the exterior ρ-ball condition with radius 1, that they have finite 1-Hausdorff dimensional measure and that the distance between the support of two different functions is precisely one. We are now going to prove that, if λ µ is small enough, then any solution of the obstacle problem (8.36) never touches the obstacles inside the domain Ω. To this aim, we first need the following lemma: Lemma Assume (8.37). Then, there exists c > 0 such that, for i = 1,..., K, we have (8.41) ψ i (x) c ν i λ µ for any x A i Ω, where ν i is the exterior normal vector to the set A i. Proof. Fix any point x 0 A i Ω. Then, by definition of A i, there exists a point z Ω c such that d(z, Ω) = µ, B λ (z) Ω A i and x 0 B λ (z). Consider now the ring {x µ < x z < λ} and the barrier function φ solution of φ = 0 in {x µ < x z < λ} φ = 1 on B µ (z) φ = 0 on B λ (z). The function ψ i is harmonic in B λ (z) Ω, ψ i 0 = φ on B λ (z) Ω and ψ i = 1 φ on Ω B λ (z). Therefore by the comparison principle, we have that ψ i (x) φ(x) for any x B λ (z) Ω, and this implies (8.41) at x = x 0. Theorem Assume (8.37). Let (u 1,..., u K ) be the limit of a converging subsequence of (u ε 1,..., uε K ), solution of (8.36). Set a := λ µ. Then, there exists a 0 > 0 such that for any a < a 0, we have, for i = 1,..., K, (8.42) u i > ψ i in A i Ω. 58

59 Proof. In order to prove (8.42), it is enough to show that (8.43) u i (x) > ψ i (x), for any x A i Ω. Indeed, if (8.43) holds true, since by (8.35) and Theorem 8.14, both u i and ψ i are harmonic in A i, the strong maximum principle implies u i > ψ i in A i. This and (8.43) give (8.42). Suppose by contradiction that there exists a point x 0 A i Ω such that u i (x 0 ) = ψ i (x 0 ) = 0. Then, by (8.41), we have that (8.44) u i (x 0 ) ψ i (x 0 ) c ν i ν i λ µ = c a. Assumptions (8.31) imply that if the angles αl i of A i at yl i, l = 1, 2, are small enough, the sets defined by and are compactly supported in Ω and Σ i = {y : y = x + ν i (x), x A i Ω} Σ i = {y : y = x + tν i (x), x A i Ω, 0 < t < 1} (8.45) d ρ (x 0, supp ψ j ) > 1 for any j i. Therefore, by (8.34), we can choose a so small that (8.45) holds true. Moreover, from (8.45), there exists a small σ > 0 such that B 1+σ (x 0 ) supp ψ j =, j i, and from (8.36), we know that u ε j 1 ε 2 uε jh(u ε i ) in B 1+σ (x 0 ) (consider u ε j extended by zero if the ball falls out of Ω). When H is defined as in (2.5) with p = 1, arguing as in (8.27) in proof of Theorem 8.10 we obtain that u j (D σ(x 0 )) 1 j i D σ(x 0 ) Now, since u i ψ i > 0 in A i and u i (x 0 ) = 0, the point x 0 belongs to {u i > 0} A i Ω. Since A i Ω has an interior tangent ball and {u i > 0} has a exterior tangent ball, we know u i. 59

60 that x 0 is a regular point. Since the set of regular points is an open set, see Lemma 8.9, for σ small enough we have (8.46) D σ(x 0 ) u i u i dh, {u i >0} D σ(x 0 ) ν i where ν i is still the exterior normal vector to A i. On another hand, if y 0 is the point that realizes the distance one with x 0, assume w.l.o.g. that y 0 supp u j, y 0 has to be in Σ i and y 0 has to be a regular point. Then, for ρ small enough such that {u j > 0} B ρ (y 0 ) is C 1 we have B ρ(y 0 ) u j u j = dh. {u j >0} B ρ(y 0 ) ν j Now, using the fact that for σ small enough such that ρ > cσ, supp u j (D σ (x 0 )) 1 B cσ (y 0 ), we have (8.47) u j u i. B cσ(y 0 ) (D σ(x 0 )) 1 Putting all together, dividing (8.46) and (8.47) respectively by H( {u i > 0} D σ (x 0 )) and H( {u j > 0} B cσ (y 0 )), and passing to the limit when σ 0 we obtain (8.48) u j ν j (y 0 ) c u i ν i (x 0 ) c a. We are now going to show that (8.48) yields a contradiction. Indeed, the point y 0 realizes its distance from the set {u i > 0} at x 0, therefore the ball B 1 (y 0 ) is tangent to {u i > 0} at x 0. Moreover, since A i {u i > 0}, the ball B 1 (y 0 ) is tangent to A i at x 0. On the other hand, for a small enough, by assumption (8.31), B 1 (y 0 ) is contained in Ω. In particular, the ρ-distance of y 0 from Ω is greater than 1. Therefore, from (8.40), we infer that u j (y 0 ) C 0, which is in contradiction with (8.48) for a small enough. When H is defined as in (2.6), we argue as in case (b) in the proof of Theorem 7.1 and similarly, we get a contradiction for a small enough. 60

61 Corollary Under the assumptions of Theorem 8.16, if a < a 0 then (u ε 1,..., uε K ) is solution of the following problem u ε i ψ i in Ω, (8.49) u ε i (x) = 1 ε 2 uε i (x) H(u ε j)(x) in Ω j i u ε i = f i on ( Ω) 1. In particular, (u ε 1,..., uε K ) is solution of (2.4). We are now ready to show that free boundaries are Lipschitz. Theorem Let (u ε 1,..., uε K ) be the solution of (2.4) given by Corollary Let (u 1,..., u K ) be the limit as ε 0 of a converging subsequence, then the free boundaries {u i > 0}, i = 1,..., K, are Lipschitz curves of the plane. Proof. By contradiction let s assume that the free boundaries are not Lipschitz. This would imply that there exists at least one singular point with asymptotic cone with zero opening. Let x 0 be an interior singular point with asymptotic cone with zero angle. W.l.o.g. suppose x 0 {u 1 > 0}. Let e 1 be the line perpendicular to the cone axis and passing through x 0, in which we choose an orientation such that the cone is below the axis e 1. As we proved in Theorem 8.10 and Corollary 8.12 there exist y 0 and y 1, with y 0, y 1 j 1 {u j > 0} singular points at distance one from x 0 with asymptotic cones with zero opening. Also, by Theorem 7.1 for any regular point x {u 1 > 0} B 1 (x 0 ) there exists a correspondent y j 1 {u j > 0} such that y = x + ν(x) with ν(x) the external normal vector to {u 1 > 0} at x. Observe that y 0, y 1 must lie on e 1. In fact, let x l n {u 1 > 0} be regular points converging to x 0, x l n x 0 as n +, from the left side of the cone axis and let x r n {u 1 > 0} be the regular points such that x r n x 0 as n +, from the right side of the cone axis. Then, the limit of the normal vectors ν(x l n) ν l and ν(x r n) ν r, are both on the direction e 1 since they are orthogonal to the cone axis. Let 61

62 y 0 and y 1 be w.l.o.g. the points defined by y 0 = x 0 + ν l y 1 = x 0 + ν r. So we have to have three singular points at distance one, all on the line e 1. Repeating the same argument and using y 1 as the reference singular point now, we conclude that there must exist another singular point, y 2, with 0 opening cone, at distance one from y 1 and also on the axis e 1. Iterating, we will be able to proceed until the prescribed boundary of the domain stops us from finding the next point. We will have all singular points with cone with zero opening aligned on the axis e 1, until we reach the boundary Ω and we cannot proceed with this process, i.e., until we cannot obtain the next point aligned in the direction of e 1 which implies that Ω crosses the axis e 1 and the distance of y k to the boundary of Ω along e 1 is less or equal than 1. Now, there are two cases: either y k Ω or y k Ω. If y k Ω assume w.l.o.g. that y k {u 1 > 0}. Since u 1 ψ 1 we have A 1 {u 1 > 0} and that y k must coincide with one of the points y 1 l, l = 1, 2, endpoints of the curve Γ 1. Indeed, by the forth assumption in (8.32), no points of {u 1 > 0} are on Ω between the curves Γ 1 and Γ 2, and Γ 1 and Γ K. Assume w.l.o.g. that y k = y1 1. Let θ be the angle of {u 1 > 0} at y1 1. Then, from (8.14) of Theorem 8.10 applied to y k = y 1 1 and y 0 = y k 1, we get θ = 0. On the other hand, since A 1 {u 1 > 0} then θ α1 1 > 0, where α1 1 is the angle of A 1 at y1 1. We have obtained a contradiction. Suppose now that y k is an interior point. Again, assume w.l.o.g. that y k {u 1 > 0}. Let z k Ω be the closest point to y k in the direction e 1 and d(y k, z k ) = l < 1. Recall that by (8.31) there is an exterior tangent ball at z k, B 1+η, so once the axis e 1 is crossed, Ω will remain outside of the tangent ball at z k and so Ω will not cross again e 1 in B 1 (y k ). We know that z k cannot belong to {u j > 0} since it does not respect the distance one and also A j {u j > 0}. And by Theorem 7.1 for any point on the free boundary there exists a correspondent point at distance one belonging to the support of another function. Taking in account the previous case, the only 62

63 f 2 y 0 x 0 y 1 l r y k 1 y k = y1 1 e f 1 1 (x l n) (x r n) B 1 (y k ) A u 1 > Figure 6. Contradiction in the case y k B 1+ y 0 x y 1 y k l 0 r z k {z} ȳ e 1 1 (x l n) (x r n) B 1 (y k ) B 1 (ȳ) u 1 > 0 Figure 7. Contradiction in the case y k Ω 63

64 option is that the point that realizes the distance from y k, ȳ, belongs to B 1 (y k ) and it must be such that the angle between e 1 and the line that contains both y k and ȳ is strictly positive, see Figure 7. Therefore, we must conclude that B 1 (ȳ) {u 1 > 0}. We have obtained a contradiction. We conclude that the free boundaries cannot have a zero angle at a singular point, therefore they are Lispschitz curves of the plane. 9. A relation between the normal derivatives at the free boundary In this section we restrict ourself to the following case: K = 2 (9.1) H defined like in (2.5), with p = 1, ϕ 1 and ρ the Euclidian norm. Therefore, the system (2.4) becomes u ε 1(x) = 1 ε 2 uε 1(x) u ε 2(x) = 1 ε 2 uε 2(x) B 1 (x) B 1 (x) u ε 2(y) dy in Ω, u ε 1(y) dy in Ω, where we denote by B 1 (x) the Euclidian ball of radius 1 centered at x. Let (u 1, u 2 ) be the limit functions of a converging subsequence that we still denote (u ε 1, uε 2 ) and for i = 1, 2 let S i := {u i > 0}. From Section 7 we know that the u i s have disjoint support and that there is a strip of width exactly one that separates S 1 and S 2. Moreover, Corollary 6.2 guarantees that at any point of the boundary of the two sets, the principal curvatures are less or equal 1. For i = 1, 2, let x i S i be such that x 1 is at distance 1 from x 2, S i is of class C 2 in a neighborhood of x i, and all the principal curvatures of S i at x i are strictly less than 1. Without loss of generality we can assume x 1 = 0 and x 2 = e n, where e n = (0,..., 1). Let us denote by u 1 ν(0) and u 2 ν(e n ) the exterior normal derivatives of u 1 and u 2 respectively at 0 and e n. Note that the two normals have opposite direction. We want to deduce a relation between u 1 ν(0) and u 2 ν(e n ). Let us start by recalling some basic properties about the level surfaces of the distance function to a set. 64

65 9.1. Level surfaces of the distance function to a set. Some basic Properties. Consider a bounded open set S and its boundary S, of the class C 2. Let κ i (x) be the principal curvatures of S at x (outward is the positive direction). Assume that for any point x S there exists a tangent ball B R (z) to S at x such that B R (z) S c. In particular the principle curvatures satisfy κ i (x) 1/R, i = 1,..., n 1. Then: a) the distance function to S, d S (x) = d(x, S), is defined and is C 2 as long as 0 < d S (x) < R. In the following lemma, which may be known in the literature, we provide a proof of the C 1,1 -regularity for a more general set, which is not necessary C 2, it may have edges as well, but it has the property that for any tangent ball there exists a clean area, in the sense explained below. For the C 2 -regularity in the case of C 2 -boundaries, see for instance Theorem in [23]. Given a bounded closed set F, we say that Π is a supporting hyperplane at x F, if x Π and there exists a ball B F c such that B is tangent to Π at x. Lemma 9.1. Let F be a bounded closed set. Assume that there exists R > 0 such that, for any x F and any supporting hyperplane Π at x, there is a ball B R (z) tangent to Π at x such that B R (z) F c. Let us denote by d F (x) = d(x, F ) the distance function from F. Then d F is of class C 1,1 in the set {0 < d F < R}. Proof. Let y 0 {0 < d F < R}. To prove that d F is of class C 1,1 at y 0, we show that there are smooth functions whose graphs are tangent from below and above the graph of d F at (y 0, d F (y 0 )). As proven in Lemma (6.3), the distance function from a closed bounded set has always a smooth tangent function from above. Indeed, let x F be a point where y 0 realizes the distance from F. Assume, without loss of generality, that x = 0. Then d(y 0, 0) = y 0 = d F (y 0 ). Moreover, the ball B y0 (y 0 ) is contained in F c and tangent to F at 0. For any y B y0 (y 0 ), we have that d F (y) d(y, 0) = y. 65

66 Therefore the cone, graph of the function y y (which is smooth at y 0 0) is tangent from above to the graph of d F at (y 0, d F (y 0 )). Next, we prove the existence of a smooth function tangent from below. Note that the tangent line to B y0 (y 0 ) at 0 is a supporting hyperplane to F at 0. Therefore, there exists a ball B R (z) tangent to F at 0 such that B R (z) F c. We must have z = Ry 0 / y 0. ) ( ) Moreover, since B R (R y 0 y 0 F c, for any y B R R y 0 y 0 {0 < d F < R}, we have that ( d F (y) d (y, B R R y 0 y 0 ( and d F (y 0 ) = y 0 = R d ) y R d ( y, R y 0 y 0 that d F is C 1,1 at y 0. y 0, R y 0 y 0 )) ( = R d y, R y ) 0 y 0 ). That is to say, the cone, graph of the function is tangent by below to the graph of d F at (y 0, d F (y 0 )). We conclude Let S(k) denote the surface that is at distance k from S S(k) := {x : d S (x) = k}, then, for k < 1 + ε and x S(k), there is a unique point x 0 S(0), such that x = x 0 + kν(x 0 ) where ν(x 0 ) is the unit normal vector at x 0 in the positive direction. More precisely, if we denote K := max{ κ i (x) : 1 i n 1, x S} and f(x, t) := x + tν(x), then f is a diffeomorphism between S ( k, k) and the neighborhood of S, N k (S) = {x + tν(x) : x S, t < k} with k < 1 K. b) for all x 0 S if we consider the linear transformation x t = x 0 + tν(x 0 ) we obtain S(t). Hence, since the tangent plane for each S(t) is always perpendicular to ν(x 0 ), the eigenvectors of the principal curvatures remain constant along the trajectories of d S, for d S < 1 + ε. c) the curvatures of S(k) satisfy, see Figure 8 1 κ i (x 0 + kν(x 0 )) = 1 κ i (x 0 ) k = κ i(x 0 ), i = 1,..., n 1, k < 1 + ε 1 κ i (x 0 )k 66

67 S(k) S } k { 0 R = 1 χ(0) Figure 8. Curvatures relation for x 0 S. d) for x 0 S, the ball B 1 (x 0 ) touches S(1) at the point x 0 +ν(x 0 ), where ν is the outward normal. Moreover, it separates quadratically from S(1), that is, for any small r > 0 and for any x B r (x 0 + ν(x 0 )) B 1 (x 0 ), we have that d(x, S(1)) Cr 2, for some C > Free boundary condition. Following Subsection 9.1, we denote by κ i (0) the principal curvatures of S 1 at 0 where outward is the positive direction and by κ i (e n ) = κ i(0) 1 κ i (0), the principal curvatures of S 2 at e n. Remark that since the normal vectors to S 1 and S 2 respectively at 0 and e n, have opposite directions, for κ i (e n ) the inner direction of S 2 is the positive one. The main result of this section is the following: Theorem 9.2. Assume (9.1). Let 0 S 1 and e n S 2. Assume that S 1 is of class C 2 in B 4h0 (0) and that the principal curvatures satisfy: κ i (0) < 1 for any i = 1,..., n 1. Then, we have the following relation: and u 1 ν(0) u 2 ν(e n ) = n 1 i=1 κ i (0) 0 κ i (0) κ i (e n ) if κ i (0) 0 for some i = 1,..., n 1, u 1 ν(0) = u 2 ν(e n ) if κ i (0) = 0 for any i = 1,..., n 1. In order to prove Theorem 9.2, we first prove a lemma that relates the mass of the Laplacians of the limit functions across the interfaces. For a point x belonging to a neighborhood of S 1 67

68 around 0, let us denote by ν(x) = ν(x 0 ) the exterior normal vector at x 0 S 1, where x 0 is the unique point such that x = x 0 + tν(x 0 ), for some small t > 0. From (a) in Subsection 9.1, ν(x) is well defined. Lemma 9.3. Under the assumptions of Theorem 9.2, for small h < h 0, let D h := B h (0) {x : d(x, S 1 ) h 2 } and E h := {y R n y = x + ν(x), x D h }. Then u 1 = u 2. D h E h Proof. Remark that the surface E h S 2 is of class C 2 for h small enough, being κ i (0) < 1 for i = 1,..., n 1, see Subsection 9.1. The Laplacians of the u i s are positive measures and u 1 = lim u ε 1 D ε 0 1(x) dx = lim h D ε 0 h ε 2 u ε 1(x)u ε 2(y) dydx, and Let s be such that ε 1 4α following way where Similarly u 2 = lim u ε 1 E ε 0 2(y) dy = lim h E ε 0 h ε 2 D h E h B 1 (x) B 1 (y) u ε 1(x)u ε 2(y) dxdy. < s < h, where α is given by Lemma 5.3. We split the set D h in the D h = D + h,s D h,s D h,s, D + h,s := {x D h d(x, S 1 ) > s 2 and u 1 (x) > 0}, D h,s := {x D h d(x, S 1 ) > s 2 and u 1 (x) = 0}, D h,s := {x D h d(x, S 1 ) s 2 }. E h = E + h,s E h,s E h,s, 68

69 where E + h,s := {x E h d(x, S 2 ) > s 2 and u 2 (x) > 0}, E h,s := {x E h d(x, S 2 ) > s 2 and u 2 (x) = 0}, E h,s := {x E h d(x, S 2 ) s 2 }, see Figure 9. Since S 1 is a smooth surface around 0, and u 1 = 0 in S 1, we have that u 1 grows linearly away from the boundary in a neighborhood of 0. This and the uniform convergence of u ε 1 to u 1, imply that there exists c > 0 such that u ε 1 (x) > cs2, for any x D + h,s for ε small enough. Then, by Lemma 5.3, u ε 2 (y) b(cs2 α ) ae ε, (a, b positive constants), for y B 1 (x) and any x D + h,s. In an analogous way, if y E+ h,s, we know that for ε small enough uε 2 (y) > cs2 and by Lemma 5.3, u ε 1 (x) b(cs2 α ) ae ε for x B 1 (y). Since we have chosen s such that s 2α > ε 1 2, we have that u ε 2 (y) = o(ε2 ) uniformly in y, for any y x D + B 1 (x) and u ε 1 (x) = o(ε2 ) uniformly h,s in x, for any x y E + B 1 (y). Remark that h,s D h,s y E + h,sb 1 (y). Therefore we have 1 ε 2 (9.2) x D h y B 1 (x) u ε 1(x)u ε 2(y)dydx = 1 ε ε ε 2 = 1 ε 2 x D + h,s x D h,s x D h,s x D h,s y B 1 (x) y B 1 (x) u ε 1(x) u ε } 2(y) dydx {{} negligible u ε 1(x)u ε 2(y)dydx u ε 1(x) }{{} y B 1 (x) negligible y B 1 (x) u ε 2(y)dydx u ε 1(x)u ε 2(y)dydx + o(1). Analogously 1 (9.3) ε 2 u ε 1(x)u ε 2(y) dxdy = 1 y E h x B 1 (y) ε 2 u ε 1(x)u ε 2(y) dxdy + o(1). E h,s B 1 (y) Next, for fixed x D h,s,we have B 1 (x) {y d(y, S 2 ) > s 2 } B 1+h (0) {y d(y, S 2 ) > s 2 } {u 2 0}. 69

70 u 2 > 0 2s E h E + h,s S 2 e n 1 2s D h,s D h 0 D h,s D + h,s S 1 u 1 > 0 Figure 9. Relation between the mass of the Laplacians Therefore for any y B 1 (x) {y d(y, S 2 ) > s 2 }, the ball B 1 (y) enters in S 1 B 2h (0) at distance at least s 2 from S 1. Since S 1 B 4h (0) is of class C 2, u 1 has linear growth away from the boundary in S 1 B 2h (0) and therefore there exists a point in B 1 (y) where u 1 cs 2 for some c > 0. Like before, Lemma 5.3 implies that u ε 2 (y) = o(ε2 ). We infer that (9.4) 1 ε 2 x D h,s y B 1 (x) u ε 1(x)u ε 2(y)dydx = 1 ε 2 x D h,s y B 1 (x) {y d(y, S 2 ) s 2 } u ε 1(x)u ε 2(y)dydx + o(1). Finally, remark that (d) of Subsection 9.1 implies that for x D h,s (9.5) B 1 (x) {y d(y, S 2 ) s 2 } E h+cs,s for some c > 0. From (9.2), (9.3), (9.4) and (9.5), we get u ε 1(x)dx = 1 D h ε 2 = 1 ε 2 1 ε 2 1 ε 2 x D h x D h,s x D h,s y B 1 (x) y E h+cs,s u ε 1(x)u ε 2(y)dydx y B 1 (x) {y d(y, S 2 ) s 2 } u ε 1(x)u ε 2(y)dydx + o(1) y E h+cs,s 70 u ε 1(x)u ε 2(y)dxdy + o(1) x B 1 (y) = u ε 2(y)dy + o(1). E h+cs u ε 1(x)u ε 2(y)dydx + o(1)

Laplace s Equation. Chapter Mean Value Formulas

Laplace s Equation. Chapter Mean Value Formulas Chapter 1 Laplace s Equation Let be an open set in R n. A function u C 2 () is called harmonic in if it satisfies Laplace s equation n (1.1) u := D ii u = 0 in. i=1 A function u C 2 () is called subharmonic

More information

PHASE TRANSITIONS: REGULARITY OF FLAT LEVEL SETS

PHASE TRANSITIONS: REGULARITY OF FLAT LEVEL SETS PHASE TRANSITIONS: REGULARITY OF FLAT LEVEL SETS OVIDIU SAVIN Abstract. We consider local minimizers of the Ginzburg-Landau energy functional 2 u 2 + 4 ( u2 ) 2 dx and prove that, if the level set is included

More information

VISCOSITY SOLUTIONS. We follow Han and Lin, Elliptic Partial Differential Equations, 5.

VISCOSITY SOLUTIONS. We follow Han and Lin, Elliptic Partial Differential Equations, 5. VISCOSITY SOLUTIONS PETER HINTZ We follow Han and Lin, Elliptic Partial Differential Equations, 5. 1. Motivation Throughout, we will assume that Ω R n is a bounded and connected domain and that a ij C(Ω)

More information

Centre for Mathematics and Its Applications The Australian National University Canberra, ACT 0200 Australia. 1. Introduction

Centre for Mathematics and Its Applications The Australian National University Canberra, ACT 0200 Australia. 1. Introduction ON LOCALLY CONVEX HYPERSURFACES WITH BOUNDARY Neil S. Trudinger Xu-Jia Wang Centre for Mathematics and Its Applications The Australian National University Canberra, ACT 0200 Australia Abstract. In this

More information

Regularity of flat level sets in phase transitions

Regularity of flat level sets in phase transitions Annals of Mathematics, 69 (2009), 4 78 Regularity of flat level sets in phase transitions By Ovidiu Savin Abstract We consider local minimizers of the Ginzburg-Landau energy functional 2 u 2 + 4 ( u2 )

More information

1. Introduction Boundary estimates for the second derivatives of the solution to the Dirichlet problem for the Monge-Ampere equation

1. Introduction Boundary estimates for the second derivatives of the solution to the Dirichlet problem for the Monge-Ampere equation POINTWISE C 2,α ESTIMATES AT THE BOUNDARY FOR THE MONGE-AMPERE EQUATION O. SAVIN Abstract. We prove a localization property of boundary sections for solutions to the Monge-Ampere equation. As a consequence

More information

Some lecture notes for Math 6050E: PDEs, Fall 2016

Some lecture notes for Math 6050E: PDEs, Fall 2016 Some lecture notes for Math 65E: PDEs, Fall 216 Tianling Jin December 1, 216 1 Variational methods We discuss an example of the use of variational methods in obtaining existence of solutions. Theorem 1.1.

More information

MINIMAL SURFACES AND MINIMIZERS OF THE GINZBURG-LANDAU ENERGY

MINIMAL SURFACES AND MINIMIZERS OF THE GINZBURG-LANDAU ENERGY MINIMAL SURFACES AND MINIMIZERS OF THE GINZBURG-LANDAU ENERGY O. SAVIN 1. Introduction In this expository article we describe various properties in parallel for minimal surfaces and minimizers of the Ginzburg-Landau

More information

Obstacle Problems Involving The Fractional Laplacian

Obstacle Problems Involving The Fractional Laplacian Obstacle Problems Involving The Fractional Laplacian Donatella Danielli and Sandro Salsa January 27, 2017 1 Introduction Obstacle problems involving a fractional power of the Laplace operator appear in

More information

A LOCALIZATION PROPERTY AT THE BOUNDARY FOR MONGE-AMPERE EQUATION

A LOCALIZATION PROPERTY AT THE BOUNDARY FOR MONGE-AMPERE EQUATION A LOCALIZATION PROPERTY AT THE BOUNDARY FOR MONGE-AMPERE EQUATION O. SAVIN. Introduction In this paper we study the geometry of the sections for solutions to the Monge- Ampere equation det D 2 u = f, u

More information

VISCOSITY SOLUTIONS OF ELLIPTIC EQUATIONS

VISCOSITY SOLUTIONS OF ELLIPTIC EQUATIONS VISCOSITY SOLUTIONS OF ELLIPTIC EQUATIONS LUIS SILVESTRE These are the notes from the summer course given in the Second Chicago Summer School In Analysis, in June 2015. We introduce the notion of viscosity

More information

THE TWO-PHASE MEMBRANE PROBLEM REGULARITY OF THE FREE BOUNDARIES IN HIGHER DIMENSIONS. 1. Introduction

THE TWO-PHASE MEMBRANE PROBLEM REGULARITY OF THE FREE BOUNDARIES IN HIGHER DIMENSIONS. 1. Introduction THE TWO-PHASE MEMBRANE PROBLEM REGULARITY OF THE FREE BOUNDARIES IN HIGHER DIMENSIONS HENRIK SHAHGHOLIAN, NINA URALTSEVA, AND GEORG S. WEISS Abstract. For the two-phase membrane problem u = λ + χ {u>0}

More information

arxiv: v2 [math.ap] 28 Nov 2016

arxiv: v2 [math.ap] 28 Nov 2016 ONE-DIMENSIONAL SAIONARY MEAN-FIELD GAMES WIH LOCAL COUPLING DIOGO A. GOMES, LEVON NURBEKYAN, AND MARIANA PRAZERES arxiv:1611.8161v [math.ap] 8 Nov 16 Abstract. A standard assumption in mean-field game

More information

ξ,i = x nx i x 3 + δ ni + x n x = 0. x Dξ = x i ξ,i = x nx i x i x 3 Du = λ x λ 2 xh + x λ h Dξ,

ξ,i = x nx i x 3 + δ ni + x n x = 0. x Dξ = x i ξ,i = x nx i x i x 3 Du = λ x λ 2 xh + x λ h Dξ, 1 PDE, HW 3 solutions Problem 1. No. If a sequence of harmonic polynomials on [ 1,1] n converges uniformly to a limit f then f is harmonic. Problem 2. By definition U r U for every r >. Suppose w is a

More information

REGULARITY FOR INFINITY HARMONIC FUNCTIONS IN TWO DIMENSIONS

REGULARITY FOR INFINITY HARMONIC FUNCTIONS IN TWO DIMENSIONS C,α REGULARITY FOR INFINITY HARMONIC FUNCTIONS IN TWO DIMENSIONS LAWRENCE C. EVANS AND OVIDIU SAVIN Abstract. We propose a new method for showing C,α regularity for solutions of the infinity Laplacian

More information

An introduction to Mathematical Theory of Control

An introduction to Mathematical Theory of Control An introduction to Mathematical Theory of Control Vasile Staicu University of Aveiro UNICA, May 2018 Vasile Staicu (University of Aveiro) An introduction to Mathematical Theory of Control UNICA, May 2018

More information

arxiv: v1 [math.ap] 25 Jul 2012

arxiv: v1 [math.ap] 25 Jul 2012 THE DIRICHLET PROBLEM FOR THE FRACTIONAL LAPLACIAN: REGULARITY UP TO THE BOUNDARY XAVIER ROS-OTON AND JOAQUIM SERRA arxiv:1207.5985v1 [math.ap] 25 Jul 2012 Abstract. We study the regularity up to the boundary

More information

Convexity in R n. The following lemma will be needed in a while. Lemma 1 Let x E, u R n. If τ I(x, u), τ 0, define. f(x + τu) f(x). τ.

Convexity in R n. The following lemma will be needed in a while. Lemma 1 Let x E, u R n. If τ I(x, u), τ 0, define. f(x + τu) f(x). τ. Convexity in R n Let E be a convex subset of R n. A function f : E (, ] is convex iff f(tx + (1 t)y) (1 t)f(x) + tf(y) x, y E, t [0, 1]. A similar definition holds in any vector space. A topology is needed

More information

JUHA KINNUNEN. Harmonic Analysis

JUHA KINNUNEN. Harmonic Analysis JUHA KINNUNEN Harmonic Analysis Department of Mathematics and Systems Analysis, Aalto University 27 Contents Calderón-Zygmund decomposition. Dyadic subcubes of a cube.........................2 Dyadic cubes

More information

Frequency functions, monotonicity formulas, and the thin obstacle problem

Frequency functions, monotonicity formulas, and the thin obstacle problem Frequency functions, monotonicity formulas, and the thin obstacle problem IMA - University of Minnesota March 4, 2013 Thank you for the invitation! In this talk we will present an overview of the parabolic

More information

Homogenization and error estimates of free boundary velocities in periodic media

Homogenization and error estimates of free boundary velocities in periodic media Homogenization and error estimates of free boundary velocities in periodic media Inwon C. Kim October 7, 2011 Abstract In this note I describe a recent result ([14]-[15]) on homogenization and error estimates

More information

TD M1 EDP 2018 no 2 Elliptic equations: regularity, maximum principle

TD M1 EDP 2018 no 2 Elliptic equations: regularity, maximum principle TD M EDP 08 no Elliptic equations: regularity, maximum principle Estimates in the sup-norm I Let be an open bounded subset of R d of class C. Let A = (a ij ) be a symmetric matrix of functions of class

More information

ESTIMATES FOR THE MONGE-AMPERE EQUATION

ESTIMATES FOR THE MONGE-AMPERE EQUATION GLOBAL W 2,p ESTIMATES FOR THE MONGE-AMPERE EQUATION O. SAVIN Abstract. We use a localization property of boundary sections for solutions to the Monge-Ampere equation obtain global W 2,p estimates under

More information

Sobolev Spaces. Chapter 10

Sobolev Spaces. Chapter 10 Chapter 1 Sobolev Spaces We now define spaces H 1,p (R n ), known as Sobolev spaces. For u to belong to H 1,p (R n ), we require that u L p (R n ) and that u have weak derivatives of first order in L p

More information

arxiv: v1 [math.ap] 10 Apr 2013

arxiv: v1 [math.ap] 10 Apr 2013 QUASI-STATIC EVOLUTION AND CONGESTED CROWD TRANSPORT DAMON ALEXANDER, INWON KIM, AND YAO YAO arxiv:1304.3072v1 [math.ap] 10 Apr 2013 Abstract. We consider the relationship between Hele-Shaw evolution with

More information

Perron method for the Dirichlet problem.

Perron method for the Dirichlet problem. Introduzione alle equazioni alle derivate parziali, Laurea Magistrale in Matematica Perron method for the Dirichlet problem. We approach the question of existence of solution u C (Ω) C(Ω) of the Dirichlet

More information

E. The Hahn-Banach Theorem

E. The Hahn-Banach Theorem E. The Hahn-Banach Theorem This Appendix contains several technical results, that are extremely useful in Functional Analysis. The following terminology is useful in formulating the statements. Definitions.

More information

u( x) = g( y) ds y ( 1 ) U solves u = 0 in U; u = 0 on U. ( 3)

u( x) = g( y) ds y ( 1 ) U solves u = 0 in U; u = 0 on U. ( 3) M ath 5 2 7 Fall 2 0 0 9 L ecture 4 ( S ep. 6, 2 0 0 9 ) Properties and Estimates of Laplace s and Poisson s Equations In our last lecture we derived the formulas for the solutions of Poisson s equation

More information

MEAN CURVATURE FLOW OF ENTIRE GRAPHS EVOLVING AWAY FROM THE HEAT FLOW

MEAN CURVATURE FLOW OF ENTIRE GRAPHS EVOLVING AWAY FROM THE HEAT FLOW MEAN CURVATURE FLOW OF ENTIRE GRAPHS EVOLVING AWAY FROM THE HEAT FLOW GREGORY DRUGAN AND XUAN HIEN NGUYEN Abstract. We present two initial graphs over the entire R n, n 2 for which the mean curvature flow

More information

Boot camp - Problem set

Boot camp - Problem set Boot camp - Problem set Luis Silvestre September 29, 2017 In the summer of 2017, I led an intensive study group with four undergraduate students at the University of Chicago (Matthew Correia, David Lind,

More information

PARTIAL REGULARITY OF BRENIER SOLUTIONS OF THE MONGE-AMPÈRE EQUATION

PARTIAL REGULARITY OF BRENIER SOLUTIONS OF THE MONGE-AMPÈRE EQUATION PARTIAL REGULARITY OF BRENIER SOLUTIONS OF THE MONGE-AMPÈRE EQUATION ALESSIO FIGALLI AND YOUNG-HEON KIM Abstract. Given Ω, Λ R n two bounded open sets, and f and g two probability densities concentrated

More information

SHARP BOUNDARY TRACE INEQUALITIES. 1. Introduction

SHARP BOUNDARY TRACE INEQUALITIES. 1. Introduction SHARP BOUNDARY TRACE INEQUALITIES GILES AUCHMUTY Abstract. This paper describes sharp inequalities for the trace of Sobolev functions on the boundary of a bounded region R N. The inequalities bound (semi-)norms

More information

Robustness for a Liouville type theorem in exterior domains

Robustness for a Liouville type theorem in exterior domains Robustness for a Liouville type theorem in exterior domains Juliette Bouhours 1 arxiv:1207.0329v3 [math.ap] 24 Oct 2014 1 UPMC Univ Paris 06, UMR 7598, Laboratoire Jacques-Louis Lions, F-75005, Paris,

More information

Lecture Notes in Advanced Calculus 1 (80315) Raz Kupferman Institute of Mathematics The Hebrew University

Lecture Notes in Advanced Calculus 1 (80315) Raz Kupferman Institute of Mathematics The Hebrew University Lecture Notes in Advanced Calculus 1 (80315) Raz Kupferman Institute of Mathematics The Hebrew University February 7, 2007 2 Contents 1 Metric Spaces 1 1.1 Basic definitions...........................

More information

Math The Laplacian. 1 Green s Identities, Fundamental Solution

Math The Laplacian. 1 Green s Identities, Fundamental Solution Math. 209 The Laplacian Green s Identities, Fundamental Solution Let be a bounded open set in R n, n 2, with smooth boundary. The fact that the boundary is smooth means that at each point x the external

More information

Equilibria with a nontrivial nodal set and the dynamics of parabolic equations on symmetric domains

Equilibria with a nontrivial nodal set and the dynamics of parabolic equations on symmetric domains Equilibria with a nontrivial nodal set and the dynamics of parabolic equations on symmetric domains J. Földes Department of Mathematics, Univerité Libre de Bruxelles 1050 Brussels, Belgium P. Poláčik School

More information

EXPOSITORY NOTES ON DISTRIBUTION THEORY, FALL 2018

EXPOSITORY NOTES ON DISTRIBUTION THEORY, FALL 2018 EXPOSITORY NOTES ON DISTRIBUTION THEORY, FALL 2018 While these notes are under construction, I expect there will be many typos. The main reference for this is volume 1 of Hörmander, The analysis of liner

More information

Local semiconvexity of Kantorovich potentials on non-compact manifolds

Local semiconvexity of Kantorovich potentials on non-compact manifolds Local semiconvexity of Kantorovich potentials on non-compact manifolds Alessio Figalli, Nicola Gigli Abstract We prove that any Kantorovich potential for the cost function c = d / on a Riemannian manifold

More information

Set, functions and Euclidean space. Seungjin Han

Set, functions and Euclidean space. Seungjin Han Set, functions and Euclidean space Seungjin Han September, 2018 1 Some Basics LOGIC A is necessary for B : If B holds, then A holds. B A A B is the contraposition of B A. A is sufficient for B: If A holds,

More information

v( x) u( y) dy for any r > 0, B r ( x) Ω, or equivalently u( w) ds for any r > 0, B r ( x) Ω, or ( not really) equivalently if v exists, v 0.

v( x) u( y) dy for any r > 0, B r ( x) Ω, or equivalently u( w) ds for any r > 0, B r ( x) Ω, or ( not really) equivalently if v exists, v 0. Sep. 26 The Perron Method In this lecture we show that one can show existence of solutions using maximum principle alone.. The Perron method. Recall in the last lecture we have shown the existence of solutions

More information

The optimal partial transport problem

The optimal partial transport problem The optimal partial transport problem Alessio Figalli Abstract Given two densities f and g, we consider the problem of transporting a fraction m [0, min{ f L 1, g L 1}] of the mass of f onto g minimizing

More information

arxiv: v2 [math.ag] 24 Jun 2015

arxiv: v2 [math.ag] 24 Jun 2015 TRIANGULATIONS OF MONOTONE FAMILIES I: TWO-DIMENSIONAL FAMILIES arxiv:1402.0460v2 [math.ag] 24 Jun 2015 SAUGATA BASU, ANDREI GABRIELOV, AND NICOLAI VOROBJOV Abstract. Let K R n be a compact definable set

More information

2) Let X be a compact space. Prove that the space C(X) of continuous real-valued functions is a complete metric space.

2) Let X be a compact space. Prove that the space C(X) of continuous real-valued functions is a complete metric space. University of Bergen General Functional Analysis Problems with solutions 6 ) Prove that is unique in any normed space. Solution of ) Let us suppose that there are 2 zeros and 2. Then = + 2 = 2 + = 2. 2)

More information

REGULARITY RESULTS FOR THE EQUATION u 11 u 22 = Introduction

REGULARITY RESULTS FOR THE EQUATION u 11 u 22 = Introduction REGULARITY RESULTS FOR THE EQUATION u 11 u 22 = 1 CONNOR MOONEY AND OVIDIU SAVIN Abstract. We study the equation u 11 u 22 = 1 in R 2. Our results include an interior C 2 estimate, classical solvability

More information

Convex solutions to the mean curvature flow

Convex solutions to the mean curvature flow Annals of Mathematics 173 (2011), 1185 1239 doi: 10.4007/annals.2011.173.3.1 Convex solutions to the mean curvature flow By Xu-Jia Wang Abstract In this paper we study the classification of ancient convex

More information

arxiv: v1 [math.ca] 7 Aug 2015

arxiv: v1 [math.ca] 7 Aug 2015 THE WHITNEY EXTENSION THEOREM IN HIGH DIMENSIONS ALAN CHANG arxiv:1508.01779v1 [math.ca] 7 Aug 2015 Abstract. We prove a variant of the standard Whitney extension theorem for C m (R n ), in which the norm

More information

REGULARITY OF THE OPTIMAL SETS FOR SOME SPECTRAL FUNCTIONALS

REGULARITY OF THE OPTIMAL SETS FOR SOME SPECTRAL FUNCTIONALS REGULARITY OF THE OPTIMAL SETS FOR SOME SPECTRAL FUNCTIONALS DARIO MAZZOLENI, SUSANNA TERRACINI, BOZHIDAR VELICHKOV Abstract. In this paper we study the regularity of the optimal sets for the shape optimization

More information

The infinity-laplacian and its properties

The infinity-laplacian and its properties U.U.D.M. Project Report 2017:40 Julia Landström Examensarbete i matematik, 15 hp Handledare: Kaj Nyström Examinator: Martin Herschend December 2017 Department of Mathematics Uppsala University Department

More information

Topological properties

Topological properties CHAPTER 4 Topological properties 1. Connectedness Definitions and examples Basic properties Connected components Connected versus path connected, again 2. Compactness Definition and first examples Topological

More information

On a Fractional Monge Ampère Operator

On a Fractional Monge Ampère Operator Ann. PDE (015) 1:4 DOI 10.1007/s40818-015-0005-x On a Fractional Monge Ampère Operator Luis Caffarelli 1 Fernando Charro Received: 16 November 015 / Accepted: 19 November 015 / Published online: 18 December

More information

Regularity of the obstacle problem for a fractional power of the laplace operator

Regularity of the obstacle problem for a fractional power of the laplace operator Regularity of the obstacle problem for a fractional power of the laplace operator Luis E. Silvestre February 24, 2005 Abstract Given a function ϕ and s (0, 1), we will stu the solutions of the following

More information

Lecture No 2 Degenerate Diffusion Free boundary problems

Lecture No 2 Degenerate Diffusion Free boundary problems Lecture No 2 Degenerate Diffusion Free boundary problems Columbia University IAS summer program June, 2009 Outline We will discuss non-linear parabolic equations of slow diffusion. Our model is the porous

More information

Calculus of Variations. Final Examination

Calculus of Variations. Final Examination Université Paris-Saclay M AMS and Optimization January 18th, 018 Calculus of Variations Final Examination Duration : 3h ; all kind of paper documents (notes, books...) are authorized. The total score of

More information

Analysis in weighted spaces : preliminary version

Analysis in weighted spaces : preliminary version Analysis in weighted spaces : preliminary version Frank Pacard To cite this version: Frank Pacard. Analysis in weighted spaces : preliminary version. 3rd cycle. Téhéran (Iran, 2006, pp.75.

More information

REAL AND COMPLEX ANALYSIS

REAL AND COMPLEX ANALYSIS REAL AND COMPLE ANALYSIS Third Edition Walter Rudin Professor of Mathematics University of Wisconsin, Madison Version 1.1 No rights reserved. Any part of this work can be reproduced or transmitted in any

More information

2. Function spaces and approximation

2. Function spaces and approximation 2.1 2. Function spaces and approximation 2.1. The space of test functions. Notation and prerequisites are collected in Appendix A. Let Ω be an open subset of R n. The space C0 (Ω), consisting of the C

More information

Regularity for Poisson Equation

Regularity for Poisson Equation Regularity for Poisson Equation OcMountain Daylight Time. 4, 20 Intuitively, the solution u to the Poisson equation u= f () should have better regularity than the right hand side f. In particular one expects

More information

REGULARITY OF POTENTIAL FUNCTIONS IN OPTIMAL TRANSPORTATION. Centre for Mathematics and Its Applications The Australian National University

REGULARITY OF POTENTIAL FUNCTIONS IN OPTIMAL TRANSPORTATION. Centre for Mathematics and Its Applications The Australian National University ON STRICT CONVEXITY AND C 1 REGULARITY OF POTENTIAL FUNCTIONS IN OPTIMAL TRANSPORTATION Neil Trudinger Xu-Jia Wang Centre for Mathematics and Its Applications The Australian National University Abstract.

More information

Optimization and Optimal Control in Banach Spaces

Optimization and Optimal Control in Banach Spaces Optimization and Optimal Control in Banach Spaces Bernhard Schmitzer October 19, 2017 1 Convex non-smooth optimization with proximal operators Remark 1.1 (Motivation). Convex optimization: easier to solve,

More information

A CONVEX-CONCAVE ELLIPTIC PROBLEM WITH A PARAMETER ON THE BOUNDARY CONDITION

A CONVEX-CONCAVE ELLIPTIC PROBLEM WITH A PARAMETER ON THE BOUNDARY CONDITION A CONVEX-CONCAVE ELLIPTIC PROBLEM WITH A PARAMETER ON THE BOUNDARY CONDITION JORGE GARCÍA-MELIÁN, JULIO D. ROSSI AND JOSÉ C. SABINA DE LIS Abstract. In this paper we study existence and multiplicity of

More information

Real Analysis Problems

Real Analysis Problems Real Analysis Problems Cristian E. Gutiérrez September 14, 29 1 1 CONTINUITY 1 Continuity Problem 1.1 Let r n be the sequence of rational numbers and Prove that f(x) = 1. f is continuous on the irrationals.

More information

The one-phase Hele-Shaw Problem with singularities.

The one-phase Hele-Shaw Problem with singularities. The one-phase Hele-Shaw Problem with singularities. David Jerison and Inwon Kim Department of Mathematics, MIT May 25, 2005 Abstract In this paper we analyze viscosity solutions of the one phase Hele-

More information

S chauder Theory. x 2. = log( x 1 + x 2 ) + 1 ( x 1 + x 2 ) 2. ( 5) x 1 + x 2 x 1 + x 2. 2 = 2 x 1. x 1 x 2. 1 x 1.

S chauder Theory. x 2. = log( x 1 + x 2 ) + 1 ( x 1 + x 2 ) 2. ( 5) x 1 + x 2 x 1 + x 2. 2 = 2 x 1. x 1 x 2. 1 x 1. Sep. 1 9 Intuitively, the solution u to the Poisson equation S chauder Theory u = f 1 should have better regularity than the right hand side f. In particular one expects u to be twice more differentiable

More information

LECTURE 15: COMPLETENESS AND CONVEXITY

LECTURE 15: COMPLETENESS AND CONVEXITY LECTURE 15: COMPLETENESS AND CONVEXITY 1. The Hopf-Rinow Theorem Recall that a Riemannian manifold (M, g) is called geodesically complete if the maximal defining interval of any geodesic is R. On the other

More information

MINIMAL GRAPHS PART I: EXISTENCE OF LIPSCHITZ WEAK SOLUTIONS TO THE DIRICHLET PROBLEM WITH C 2 BOUNDARY DATA

MINIMAL GRAPHS PART I: EXISTENCE OF LIPSCHITZ WEAK SOLUTIONS TO THE DIRICHLET PROBLEM WITH C 2 BOUNDARY DATA MINIMAL GRAPHS PART I: EXISTENCE OF LIPSCHITZ WEAK SOLUTIONS TO THE DIRICHLET PROBLEM WITH C 2 BOUNDARY DATA SPENCER HUGHES In these notes we prove that for any given smooth function on the boundary of

More information

B. Appendix B. Topological vector spaces

B. Appendix B. Topological vector spaces B.1 B. Appendix B. Topological vector spaces B.1. Fréchet spaces. In this appendix we go through the definition of Fréchet spaces and their inductive limits, such as they are used for definitions of function

More information

Minimal Surface equations non-solvability strongly convex functional further regularity Consider minimal surface equation.

Minimal Surface equations non-solvability strongly convex functional further regularity Consider minimal surface equation. Lecture 7 Minimal Surface equations non-solvability strongly convex functional further regularity Consider minimal surface equation div + u = ϕ on ) = 0 in The solution is a critical point or the minimizer

More information

Regularity for the One-Phase Hele-Shaw problem from a Lipschitz initial surface

Regularity for the One-Phase Hele-Shaw problem from a Lipschitz initial surface Regularity for the One-Phase Hele-Shaw problem from a Lipschitz initial surface Sunhi Choi, David Jerison and Inwon Kim June 2, 2005 Abstract In this paper we show that if the Lipschitz constant of the

More information

NECESSARY CONDITIONS FOR WEIGHTED POINTWISE HARDY INEQUALITIES

NECESSARY CONDITIONS FOR WEIGHTED POINTWISE HARDY INEQUALITIES NECESSARY CONDITIONS FOR WEIGHTED POINTWISE HARDY INEQUALITIES JUHA LEHRBÄCK Abstract. We establish necessary conditions for domains Ω R n which admit the pointwise (p, β)-hardy inequality u(x) Cd Ω(x)

More information

Mathematics for Economists

Mathematics for Economists Mathematics for Economists Victor Filipe Sao Paulo School of Economics FGV Metric Spaces: Basic Definitions Victor Filipe (EESP/FGV) Mathematics for Economists Jan.-Feb. 2017 1 / 34 Definitions and Examples

More information

OPTIMAL DISTRIBUTION OF OPPOSITELY CHARGED PHASES: PERFECT SCREENING AND OTHER PROPERTIES

OPTIMAL DISTRIBUTION OF OPPOSITELY CHARGED PHASES: PERFECT SCREENING AND OTHER PROPERTIES OPTIMAL DISTRIBUTION OF OPPOSITELY CHARGED PHASES: PERFECT SCREENING AND OTHER PROPERTIES MARCO BONACINI, HANS KNÜPFER, AND MATTHIAS RÖGER Abstract. We study the minimum energy configuration of a uniform

More information

In this chapter we study elliptical PDEs. That is, PDEs of the form. 2 u = lots,

In this chapter we study elliptical PDEs. That is, PDEs of the form. 2 u = lots, Chapter 8 Elliptic PDEs In this chapter we study elliptical PDEs. That is, PDEs of the form 2 u = lots, where lots means lower-order terms (u x, u y,..., u, f). Here are some ways to think about the physical

More information

A NONLINEAR OPTIMIZATION PROBLEM IN HEAT CONDUCTION

A NONLINEAR OPTIMIZATION PROBLEM IN HEAT CONDUCTION A NONLINEAR OPTIMIZATION PROBLEM IN HEAT CONDUCTION EDUARDO V. TEIXEIRA Abstract. In this paper we study the existence and geometric properties of an optimal configuration to a nonlinear optimization problem

More information

Course 212: Academic Year Section 1: Metric Spaces

Course 212: Academic Year Section 1: Metric Spaces Course 212: Academic Year 1991-2 Section 1: Metric Spaces D. R. Wilkins Contents 1 Metric Spaces 3 1.1 Distance Functions and Metric Spaces............. 3 1.2 Convergence and Continuity in Metric Spaces.........

More information

EXISTENCE RESULTS FOR QUASILINEAR HEMIVARIATIONAL INEQUALITIES AT RESONANCE. Leszek Gasiński

EXISTENCE RESULTS FOR QUASILINEAR HEMIVARIATIONAL INEQUALITIES AT RESONANCE. Leszek Gasiński DISCRETE AND CONTINUOUS Website: www.aimsciences.org DYNAMICAL SYSTEMS SUPPLEMENT 2007 pp. 409 418 EXISTENCE RESULTS FOR QUASILINEAR HEMIVARIATIONAL INEQUALITIES AT RESONANCE Leszek Gasiński Jagiellonian

More information

THE STEINER REARRANGEMENT IN ANY CODIMENSION

THE STEINER REARRANGEMENT IN ANY CODIMENSION THE STEINER REARRANGEMENT IN ANY CODIMENSION GIUSEPPE MARIA CAPRIANI Abstract. We analyze the Steiner rearrangement in any codimension of Sobolev and BV functions. In particular, we prove a Pólya-Szegő

More information

Harmonic Functions and Brownian motion

Harmonic Functions and Brownian motion Harmonic Functions and Brownian motion Steven P. Lalley April 25, 211 1 Dynkin s Formula Denote by W t = (W 1 t, W 2 t,..., W d t ) a standard d dimensional Wiener process on (Ω, F, P ), and let F = (F

More information

l(y j ) = 0 for all y j (1)

l(y j ) = 0 for all y j (1) Problem 1. The closed linear span of a subset {y j } of a normed vector space is defined as the intersection of all closed subspaces containing all y j and thus the smallest such subspace. 1 Show that

More information

Real Analysis Notes. Thomas Goller

Real Analysis Notes. Thomas Goller Real Analysis Notes Thomas Goller September 4, 2011 Contents 1 Abstract Measure Spaces 2 1.1 Basic Definitions........................... 2 1.2 Measurable Functions........................ 2 1.3 Integration..............................

More information

REGULARITY OF MONOTONE TRANSPORT MAPS BETWEEN UNBOUNDED DOMAINS

REGULARITY OF MONOTONE TRANSPORT MAPS BETWEEN UNBOUNDED DOMAINS REGULARITY OF MONOTONE TRANSPORT MAPS BETWEEN UNBOUNDED DOMAINS DARIO CORDERO-ERAUSQUIN AND ALESSIO FIGALLI A Luis A. Caffarelli en su 70 años, con amistad y admiración Abstract. The regularity of monotone

More information

Non-radial solutions to a bi-harmonic equation with negative exponent

Non-radial solutions to a bi-harmonic equation with negative exponent Non-radial solutions to a bi-harmonic equation with negative exponent Ali Hyder Department of Mathematics, University of British Columbia, Vancouver BC V6TZ2, Canada ali.hyder@math.ubc.ca Juncheng Wei

More information

Geometric and isoperimetric properties of sets of positive reach in E d

Geometric and isoperimetric properties of sets of positive reach in E d Geometric and isoperimetric properties of sets of positive reach in E d Andrea Colesanti and Paolo Manselli Abstract Some geometric facts concerning sets of reach R > 0 in the n dimensional Euclidean space

More information

Hartogs Theorem: separate analyticity implies joint Paul Garrett garrett/

Hartogs Theorem: separate analyticity implies joint Paul Garrett  garrett/ (February 9, 25) Hartogs Theorem: separate analyticity implies joint Paul Garrett garrett@math.umn.edu http://www.math.umn.edu/ garrett/ (The present proof of this old result roughly follows the proof

More information

Partial regularity for fully nonlinear PDE

Partial regularity for fully nonlinear PDE Partial regularity for fully nonlinear PDE Luis Silvestre University of Chicago Joint work with Scott Armstrong and Charles Smart Outline Introduction Intro Review of fully nonlinear elliptic PDE Our result

More information

Growth Theorems and Harnack Inequality for Second Order Parabolic Equations

Growth Theorems and Harnack Inequality for Second Order Parabolic Equations This is an updated version of the paper published in: Contemporary Mathematics, Volume 277, 2001, pp. 87-112. Growth Theorems and Harnack Inequality for Second Order Parabolic Equations E. Ferretti and

More information

Behavior of space periodic laminar flames near the extinction point

Behavior of space periodic laminar flames near the extinction point Behavior of space periodic laminar flames near the extinction point Sunhi Choi Abstract In this paper we study a free boundary problem, arising from a model for the propagation of laminar flames. Consider

More information

Analysis Finite and Infinite Sets The Real Numbers The Cantor Set

Analysis Finite and Infinite Sets The Real Numbers The Cantor Set Analysis Finite and Infinite Sets Definition. An initial segment is {n N n n 0 }. Definition. A finite set can be put into one-to-one correspondence with an initial segment. The empty set is also considered

More information

Math General Topology Fall 2012 Homework 6 Solutions

Math General Topology Fall 2012 Homework 6 Solutions Math 535 - General Topology Fall 202 Homework 6 Solutions Problem. Let F be the field R or C of real or complex numbers. Let n and denote by F[x, x 2,..., x n ] the set of all polynomials in n variables

More information

NOTES ON SCHAUDER ESTIMATES. r 2 x y 2

NOTES ON SCHAUDER ESTIMATES. r 2 x y 2 NOTES ON SCHAUDER ESTIMATES CRISTIAN E GUTIÉRREZ JULY 26, 2005 Lemma 1 If u f in B r y), then ux) u + r2 x y 2 B r y) B r y) f, x B r y) Proof Let gx) = ux) Br y) u r2 x y 2 Br y) f We have g = u + Br

More information

EXISTENCE OF SOLUTIONS TO ASYMPTOTICALLY PERIODIC SCHRÖDINGER EQUATIONS

EXISTENCE OF SOLUTIONS TO ASYMPTOTICALLY PERIODIC SCHRÖDINGER EQUATIONS Electronic Journal of Differential Equations, Vol. 017 (017), No. 15, pp. 1 7. ISSN: 107-6691. URL: http://ejde.math.txstate.edu or http://ejde.math.unt.edu EXISTENCE OF SOLUTIONS TO ASYMPTOTICALLY PERIODIC

More information

Oblique derivative problems for elliptic and parabolic equations, Lecture II

Oblique derivative problems for elliptic and parabolic equations, Lecture II of the for elliptic and parabolic equations, Lecture II Iowa State University July 22, 2011 of the 1 2 of the of the As a preliminary step in our further we now look at a special situation for elliptic.

More information

Sobolev spaces. May 18

Sobolev spaces. May 18 Sobolev spaces May 18 2015 1 Weak derivatives The purpose of these notes is to give a very basic introduction to Sobolev spaces. More extensive treatments can e.g. be found in the classical references

More information

Austin Mohr Math 730 Homework. f(x) = y for some x λ Λ

Austin Mohr Math 730 Homework. f(x) = y for some x λ Λ Austin Mohr Math 730 Homework In the following problems, let Λ be an indexing set and let A and B λ for λ Λ be arbitrary sets. Problem 1B1 ( ) Show A B λ = (A B λ ). λ Λ λ Λ Proof. ( ) x A B λ λ Λ x A

More information

Chapter 1. Measure Spaces. 1.1 Algebras and σ algebras of sets Notation and preliminaries

Chapter 1. Measure Spaces. 1.1 Algebras and σ algebras of sets Notation and preliminaries Chapter 1 Measure Spaces 1.1 Algebras and σ algebras of sets 1.1.1 Notation and preliminaries We shall denote by X a nonempty set, by P(X) the set of all parts (i.e., subsets) of X, and by the empty set.

More information

SYMMETRY RESULTS FOR PERTURBED PROBLEMS AND RELATED QUESTIONS. Massimo Grosi Filomena Pacella S. L. Yadava. 1. Introduction

SYMMETRY RESULTS FOR PERTURBED PROBLEMS AND RELATED QUESTIONS. Massimo Grosi Filomena Pacella S. L. Yadava. 1. Introduction Topological Methods in Nonlinear Analysis Journal of the Juliusz Schauder Center Volume 21, 2003, 211 226 SYMMETRY RESULTS FOR PERTURBED PROBLEMS AND RELATED QUESTIONS Massimo Grosi Filomena Pacella S.

More information

APPLICATIONS OF DIFFERENTIABILITY IN R n.

APPLICATIONS OF DIFFERENTIABILITY IN R n. APPLICATIONS OF DIFFERENTIABILITY IN R n. MATANIA BEN-ARTZI April 2015 Functions here are defined on a subset T R n and take values in R m, where m can be smaller, equal or greater than n. The (open) ball

More information

Regularity estimates for the solution and the free boundary of the obstacle problem for the fractional Laplacian

Regularity estimates for the solution and the free boundary of the obstacle problem for the fractional Laplacian Regularity estimates for the solution and the free boundary of the obstacle problem for the fractional Laplacian Luis Caffarelli, Sandro Salsa and Luis Silvestre October 15, 2007 Abstract We use a characterization

More information

Deforming conformal metrics with negative Bakry-Émery Ricci Tensor on manifolds with boundary

Deforming conformal metrics with negative Bakry-Émery Ricci Tensor on manifolds with boundary Deforming conformal metrics with negative Bakry-Émery Ricci Tensor on manifolds with boundary Weimin Sheng (Joint with Li-Xia Yuan) Zhejiang University IMS, NUS, 8-12 Dec 2014 1 / 50 Outline 1 Prescribing

More information

Metric Spaces. Exercises Fall 2017 Lecturer: Viveka Erlandsson. Written by M.van den Berg

Metric Spaces. Exercises Fall 2017 Lecturer: Viveka Erlandsson. Written by M.van den Berg Metric Spaces Exercises Fall 2017 Lecturer: Viveka Erlandsson Written by M.van den Berg School of Mathematics University of Bristol BS8 1TW Bristol, UK 1 Exercises. 1. Let X be a non-empty set, and suppose

More information

Threshold solutions and sharp transitions for nonautonomous parabolic equations on R N

Threshold solutions and sharp transitions for nonautonomous parabolic equations on R N Threshold solutions and sharp transitions for nonautonomous parabolic equations on R N P. Poláčik School of Mathematics, University of Minnesota Minneapolis, MN 55455 Abstract This paper is devoted to

More information