On the Yamabe problem and the scalar curvature problems under ...

Report 1 Downloads 49 Views
Math. Ann. 322, 667–699 (2002)

Mathematische Annalen

Digital Object Identifier (DOI) 10.1007/s002080100267

On the Yamabe problem and the scalar curvature problems under boundary conditions Antonio Ambrosetti · YanYan Li · Andrea Malchiodi Received: 18 May 2000 / Published online: 28 February 2002 – © Springer-Verlag 2002

1. Introduction In this paper we prove some existence results concerning a problem arising in conformal differential geometry. Consider a smooth metric g on B = {x ∈ Rn : |x| < 1}, the unit ball on Rn , n ≥ 3, and let ∆g , Rg , νg , hg denote, respectively, the Laplace-Beltrami operator, the scalar curvature of (B, g), the outward unit normal to ∂B = S n−1 with respect to g and the mean curvature of (S n−1 , g). Given two smooth functions R and h , we will be concerned with the existence of positive solutions u ∈ H 1 (B) of  (n − 1) n+2   −4 (n − 2) ∆g u + Rg u = R u n−2 , in B;   

2 n ∂νg u + hg u = h u n−2 , on ∂B = S n−1 . (n − 2)

(1)

It is well known that such a solution is C ∞ provided g, R and h are, see [10]. If u > 0 is a smooth solution of (1) then g = u4/(n−2) g is a metric, conformally equivalent to g, such that R and h are, respectively, the scalar curvature of (B, g ) and the mean curvature of (S n−1 , g ). Up to a stereographic projection, this is equivalent to finding a conformal metric on the upper half sphere S+n = {(x1 , . . . , xn+1 ) ∈ Rn+1 : |x| = 1, xn+1 > 0} such that the scalar curvature of S+n and the mean curvature of ∂S+n = S n−1 are prescribed functions. A. Ambrosetti S.I.S.S.A., via Beirut, 2–4, 34019 Trieste, Italy (e-mail: [email protected]) Y.Y. Li, A. Malchiodi Department of Mathematics, Rutgers University New Brunswick, NJ 08903, USA (e-mail: [email protected]; [email protected]) The first and the third authors have been supported by the italian M.U.R.S.T. under the national project Variational Methods and Nonlinear Differential Equations.

668

A. Ambrosetti et al.

In the first part of the paper we deal with the the case in which R and h are constant, say R ≡ 1 and h ≡ c, when (1) becomes  (n − 1) n+2   −4 (n − 2) ∆g u + Rg u = u n−2 , in B; (Y )  2 n  n−1  ∂ν u + hg u = cu n−2 , on ∂B = S . (n − 2) g This will be referred as the Yamabe like problem and was first studied in [10– 12]. More recently, the existence of a solution of (1) has been proved in [14,15] under the assumption that (B, g) is of positive type (for a definition see [14]) and satisfies one of the following assumptions: (i) (B, g) is locally conformally flat and ∂B is umbilical; (ii) n ≥ 5 and ∂B is not umbilical. Our main result concerning the Yamabe like problem shows that none of (i) or (ii) is required when g is close to the standard metric g0 on B. Precisely, consider the following class Gε of bilinear forms Gε = {g ∈ C ∞ (B) : g − g0 L∞ (B) ≤ ε, ∇gLn (B) ≤ ε, ∇gLn−1 (S n−1 ) ≤ ε}. (2) Inequalities in (2) hold if for example g−g0 C 1 (B) ≤ ε, or if g−g0 W 2,n (B) ≤ ε. We will show: Theorem 1. Given M > 0 there exists ε0 > 0 such that for every ε with ε ∈ (0, ε0 ), for every c > −M and for every metric g ∈ Gε problem (Y ) possesses a positive solution. In the second part of the paper we will take g = g0 , R = 1 + εK(x), h = c + εh(x) and consider the Scalar Curvature like problem  (n − 1) n+2   −4 (n − 2) ∆u = (1 + εK(x))u n−2 , in B; (Pε )  2 ∂u n  n−1  n−2 + u = (c + εh(x)) u , on S , (n − 2) ∂ν

where ν = νg0 . The Scalar Curvature like problem has been studied in [16] where a non perturbative problem like  (n − 1) n+2   −4 (n − 2) ∆u = R (x)u n−2 , in B;  2 ∂u   + u = 0, on S n−1 , (n − 2) ∂ν

Yamabe and scalar curvature

669

has been considered. We also mention the paper [9] dealing with the existence of solutions of    ∆u = 0, in B; (3) 2 ∂u n  n−1  n−2 + u = (1 + εh(x)) u , on S ,  (n − 2) ∂ν a problem similar in nature to (Pε ). To give an idea of the existence results we can prove, let us consider the particular cases that either h ≡ 0 or K ≡ 0. In the former, problem (Pε ) becomes  (n − 1) n+2   −4 (n − 2) ∆u = (1 + εK(x))u n−2 , in B; (Pε,K )  2 ∂u n  n−1  n−2 + u = c u , on S , (n − 2) ∂ν Theorem 2. Suppose that K satisfies (K1 ) there exists an absolute maximum (resp. minimum) p of K|S n−1 such that K (p) · p < 0, resp. K (p) · p > 0. Then for |ε| sufficiently small, (Pε,K ) has a positive solution. Another kind of result is the following Theorem 3. Let K|S n−1 be a Morse function and satisfies K (x) · x = 0,  x∈Crit (K|S n−1

∀ x ∈ Crit (K|S n−1 )

(K2 )

(−1)m(x,K) = 1,

(K3 )

):K (x)·x 0.

(5)

Z c = {zµ,ξ : µ2 + |ξ |2 − cκµ − 1 = 0}

(6)

The set

is an n-dimensional manifold, diffeomorphic to a ball in Rn , with boundary ∂Z c corresponding to the parameter values µ = 0, |ξ | = 1.

Yamabe and scalar curvature

673

We need to study the eigenvalues of I0 (zµ,ξ ), with zµ,ξ ∈ Z c . Recall that, by definition, λ ∈ R is an eigenvalue of I0 (zµ,ξ ) if there exists v ∈ H 1 (B), v = 0 such that I0 (zµ,ξ )[v] = λv and this means that v is solution of the linear problem  (n − 1) 4 n + 2 n−2  −4 − λ) ∆v = z v, (1   (n − 2) n − 2 µ,ξ

2  (n − 1) n  n−2 4 z + λ − 1 v, (1 − λ) ∂ν v = 2(n − 1) c (n − 2) (n − 2) µ,ξ

in B; on S n−1 . (7)

The following lemma is well known. Lemma 1. (a) λ = 0 is an eigenvalue of (7) and the corresponding eigenspace is n dimensional and coincides with the tangent space to Z c at zµ,ξ , namely is spanned by Dzµ,ξ . (b) (7) has precisely one negative eigenvalue λ1 (c); all the remaining eigenvalues are positive. Item (a) is proved in [14]. Item (b) easily follows from the fact that zµ,ξ is a Mountain Pass critical point of I0c . Let λ2 (c) denote the smallest positive eigenvalue of I0 (zµ,ξ ). The main result of this section is the following one: Lemma 2. For all M > 0 there exists a positive constant CM such that 1 ≤ |λi (c)| ≤ CM , ∀ c ≥ −M, i = 1, 2. CM Remark. There is a numerical evidence that λ2 (c) ↓ 0 as c ↓ −∞. Proof. We will prove separately that |λi (c)| ≤ CM and that C1M ≤ |λi (c)|. For symmetry reasons it is sufficient to take zµ,ξ = zµ , namely to take ξ = 0. In such a case µ depends only on ξ and (5) yields

1 µ(c) = κc + κ 2 c2 + 4 . 2 Case 1. |λi (c)| ≤ CM . By contradiction suppose there exists a sequence cj → +∞ such that |λi (cj )| → +∞, i = 1, 2. Let vj denote an eigenfunction of (7) with λ = λi (cj ). Then vj solves the problem  ∆vj = aj (x)vj ,

in B;

∂ v = b (x)v , ν j j j

on S n−1 ,

(8)

674

A. Ambrosetti et al.

where 4 n + 2 n−2 1 zµ(cj ) (x), x ∈ B (λi (cj ) − 1) 4(n − 1)

2 n n−2 n−2 cj z (x) + λi (cj ) − 1 , x ∈ S n−1 . bj (x) = 2(1 − λi (cj )) (n − 2) µ(cj )

aj (x) =

Above, it is worth pointing out that bj is constant on S n−1 . Actually, there results 2 n−2

zµ (x) = κ µ

−1



1 1+ 2 µ

−1

, ∀ x ∈ S n−1 ,

and hence

 

−1 n n−2 1 −1 bj ≡ cj · κ µ (cj ) 1 + 2 + λi (cj ) − 1 , 2(1 − λi (cj )) (n − 2) µ (cj ) ∀ x ∈ S n−1 .

Moreover, since µ ∼ κc as c → +∞, it turns out that bj → −

(n − 2) . 2

(9)

Now, integrating by parts we deduce from (8)    |∇vj |2 dx + aj vj2 dx = bj B

S n−1

B

vj2 dσ.

(10)

Using (9) and a Poincar´e-like inequality, we find there exists C > 0 1   2 − aj vj dx ≥ C vj2 dx. B

B

This leads to a contradiction because aj (x) → 0 in C 0 (B) and vj ≡ 0. Case 2. C1M ≤ |λi (c)|. Arguing again by contradiction, let cj → +∞ and suppose that |λi (cj )| → 0. As before, the corresponding eigenfunctions vj satisfy (10), where now bj → 1, because µ ∼ κc and |λi (cj )| → 0. Choosing vj is such a way that supB |vj | = 1, then (10) yields that vj is bounded in H 1 (B) and hence vj < v0 weakly in H 1 (B). Passing to the limit in    ∇vj · ∇w + aj vj w − bj vj w = 0, ∀w ∈ H 1 (B), B

B

S n−1

1 in the sequel we will use the same symbol C to denote possibly different positive constants.

Yamabe and scalar curvature

675

it immedately follows that v0 satisfies  ∆v0 = 0, ∂ v = v , ν 0 0

in B; on S n−1 .

(P3 )

The solutions of problem (P3 ) are explicitly known, namely they are the linear functions an B. We denote by X the vector space of these solutions, which is n-dimensional. To complete the proof we will show that v0 ∈ X leads to a contradiction. We know that λ = 0 is an eigenvalue with multiplicity n, and the eigenvectors corresponding to λ = 0 are precisely the elements of Tzµ Z c . Let uj ∈ Tzµ (cj ) Z c with supB |uj | = 1. Then, by using simple computations, one can prove that, up to a subsequence, uj → v strongly in H 1 (B) for some function v ∈ X. We can assume w.l.o.g. that v = v0 (the weak limit of vj ), so it follows that (uj , vj ) → v0 2 = 0. But this is not possible, since vj are eigenvectors corresponding to λ1 < 0, while uj are eigenvectors corresponding to λ = 0 and hence they are orthogonal.   In conclusion, taking into account of Lemma 2, we can state: Lemma 3. The unperturbed functional I0c possesses an n-dimensional manifold Z c of critical points, diffeomorphic to a ball of Rn . Moreover I0c satisfies the following properties (i) I0 (z) = I − K, where K is a compact operator for every z ∈ Z c ; (ii) Tz Z c = KerD 2 I0c (z) for all z ∈ Z c . From (i)-(ii) it follows that the restriction of D 2 I0c to (Tz Z c )⊥ is invertible. Moreover, denoting by Lc (z) its inverse, for every M > 0 there exists C > 0 such that Lc (z) ≤ C for all z ∈ Z c

and for all c > −M.

(11)

3. The Yamabe like problem 3.1. Preliminaries Solution s of problem (1) can be found as critical points of the functional I c : H 1 (B) → R defined in (4). We recall some formulas from [3] which will be useful for our computations. We denote with gij the coefficients of the metric g in some local co-ordinates and with g ij the elements of the inverse matrix (g −1 )ij .

676

A. Ambrosetti et al.

The volume element dVg of the metric g ∈ Gε , taking into account (2) is 1

dVg = |g| 2 · dx = (1 + O(ε)) · dx 2 .

(12)

The Christoffel symbols are given by Γijl = 21 [Di gkj + Dj gki − Dk gij ]g kl . The components of the Riemann tensor, the Ricci tensor and the scalar curvature are, respectively l l l = Di Γjlk − Dj Γikl + Γim Γjmk − Γjlm Γikm ; Rkj = Rklj ; R = Rg = Rkj g kj . Rkij (13)

For a smooth function u the components of ∇g u are (∇g u)i = g ij Dj u, so (∇g u)i = ∇u · (1 + O(ε)).

(14)

From the preceding formulas and from the fact that g ∈ Gε it readily follows that I c (u) = I0c (u) + O(ε). More precisely, the following lemma holds. The proof is rather technical and is postponed to the Appendix. Lemma 4. Given M > 0 there exists C > 0 such that for c > −M and g ∈ Gε there holds ∇I c (z) ≤ C · ε · (1 + |c|)−

n−2 2

, ∀z ∈ Z c ;

 2 c  D I (z) − D 2 I c (z) ≤ C · ε, ∀z ∈ Z c 0  c  I (z + w) − I c (z + w ) ≤ C · (1 + |c|) · (ε + ρ



2 n−2

(15)

(16)

(17)

) · w − w , ∀z ∈ Z c , w, w ∈ H 1 (B), ∀w, w  ≤ ρ;

4

4

1 + u n−2 + w n−2

 c  ∇I (u + w) − ∇I c (u) ≤ C · w · (18)

2 2 + |c| · u n−2 + |c| · w n−2 , ∀u, w ∈ H 1 (B).

Moreover, if u is uniformly bounded and if w ≤ 1 there results   2 c 2 D I (u + w) − D 2 I c (u) ≤ C · (1 + |c|) · w n−2 .

(19)

Yamabe and scalar curvature

677

3.2. A finite dimensional reduction The aim of this section is to perform a finite dimensional reduction, using Lemma 3. Arguments of this kind has been emploied, e.g. in [1]. The first step is to construct, for g ∈ Gε , a perturbed manifold Zgc  Z c which is a natural constraint for I c , namely: if u ∈ Zgc and ∇I c |Zgc (u) = 0 then ∇I c (u) = 0. For brevity, we denote by z˙ ∈H 1 (B))n an orthonormal n-tuple in Tz Z c . Moreover, if α ∈ Rn we set α z˙ = αi z˙ i . Proposition 1. Given M > 0, there exist ε0 , C > 0, such that ∀ c > −M, ∀ z ∈ Zc ∀ ε ≤ ε0 and ∀ g ∈ Gε there are C 1 functions w = w(z, g, c) ∈ H 1 (B) and α = α(z, g, c) ∈ Rn such that the following properties hold (i) w is orthogonal to Tz Z c ∀z ∈ Z c , i.e. (w, z˙ ) = 0; (ii) ∇I c (z + w) = α z˙ ∀z ∈ Z c ; n−2 (iii) (w, α) ≤ C · ε · (1 + |c|)− 2 ∀z ∈ Z c . Furthermore, from (i)-(ii) it follows that (iv) the manifold Zgc = {z + w(z, g, c) | z ∈ Z c } is a natural constraint for I c . Proof. Let us define 3 Hg : Z c × H 1 (B) × Rn → H 1 (B) × Rn by setting

c ∇I (z + w) − α z˙ . Hg (z, w, α) = (w, z˙ ) With this notation, the unknown (w, α) can be implicitly defined by the equation Hg (z, w, α) = (0, 0). Setting Rg (z, w, α) = Hg (z, w, α) − ∂(w,α) Hg (z, 0, 0) [(w, α)] we have that Hg (z, w, α) = 0



∂(w,α) Hg (z, 0, 0)[(w, α)] + Rg (z, w, α) = 0.

Let H0 = Hg0 . From (11) it follows easily that ∂(w,α) H0 (z, 0, 0) is invertible uniformly w.r.t. z ∈ Z c and c > −M. Moreover using (16) it turns out that for ε0 sufficiently small and for ε ≤ ε0 also the operator ∂(w,α) Hg (z, 0, 0) is invertible and has uniformly bounded inverse, provided g ∈ Gε . Hence, for such g there results Hg (z, w, α) = 0 ⇔ (w, α) = Fz,g (w, α) −1  Rg (z, w, α). := − ∂(w,α) Hg (z, 0, 0) We prove the Proposition by showing that the map Fz,g is a contraction in some ball Bρ = {(w, α) ∈ H 1 (B) × Rn : w + |α| ≤ ρ}, with ρ of order 3 H depends also on c, but such a dependence will be understood.

678

A. Ambrosetti et al. n−2

ρ ∼ ε · (1 + |c|)− 2 . We first show that there exists C > 0 such that for all (w, α), (w , α ) ∈ Bρ , all z ∈ Z c and all g ∈ Gε , there holds 

n  Fz,g (w, α) ≤ C · ε · (1 + |c|)− n−2 2 + (1 + |c|) · ρ n−2 , 2  F (w , α ) − F (w, α) ≤ C · (1 + |c|) · ρ n−2 · (w, α) − (w , α ). z,g

z,g

(20) Condition (20) is equivalent to the following two inequalities

n−2 2 ∇I c (z + w) − D 2 I c (z)[w] ≤ C · ε · (1 + |c|)− 2 + (1 + |c|) · ρ n−2 ; (21) (∇I c (z + w) − D 2 I c (z)[w]) − (∇I c (z + w ) − D 2 I c (z)[w ]) ≤ (22) C · (1 + |c|) · ρ n−2 · (w, α) − (w , α ). 2

Let us first prove (21). There holds ∇I c (z + w) − D 2 I c (z)[w] = ∇I c (z + w) − ∇I c (z) + ∇I c (z) − D 2 I c (z)[w]  1   2 c c D I (z + sw) − D 2 I c (z) [w]ds. = ∇I (z) + 0

Hence it turns out that ∇I c (z+w)−D 2 I c (z)[w] ≤ ∇I c (z)+w· sup D 2 I c (z+sw)−D 2 I c (z). s∈[0,1]

Using (19) we have n

∇I c (z + w) − D 2 I c (z)[w] ≤ ∇I c (z) + C · (1 + |c|) · ρ n−2 . Hence from (15) we deduce that



n−2 n ∇I c (z + w) − D 2 I c (z)[w] ≤ C · ε · (1 + |c|)− 2 + (1 + |c|) · ρ n−2 ,

and (21) follows. We turn now to (22). There holds ∇I c (z + w) − ∇I c (z + w ) − D 2 I c (z)[w − w ]   1

  2 c 2 c D I (z + w + s(w − w)) − D I (z) [w − w]ds  =   0

≤ sup D 2 I c (z + w + s(w − w)) − D 2 I c (z) · w − w. s∈[0,1]

Using again (19), and taking w, w  ≤ ρ we have that ||D 2 I c (z + w + s(w − w )) − D 2 I c (z)|| ≤ C · (1 + |c|) · ρ n−2 , 2

Yamabe and scalar curvature

679 n−2

proving (22). Taking ρ = 2C · ε · (1 + |c|)− 2 and ε ≤ ε0 , with ε0 sufficiently small, there results 

n C · ε · (1 + |c|)− n−2 2 + (1 + |c|) · ρ n−2 < ρ, 2 C · (1 + |c|) · ρ n−2 < 1.

Then Fz,g is a contraction in Bρ and hence Hg = 0 has a unique solution w = n−2   w(z, g, c), α = α(z, g, c) with (w, α) ≤ 2C · ε · (1 + |c|)− 2 . Remark 1. In general, the preceding arguments give rise to the following result, see [1]. Let Iε (u) = I0 (u) + O(ε) denote a C 2 functional and suppose that I0 has an n-dimensional manifold Z of critical points satisfying (i) − (ii) of Lemma 3. Then for |ε| small there exists a unique w = wε (z) satisfying (i) − (ii) − (iii) of Proposition 1. Furthermore, the manifold Zε = {z + wε (z) : z ∈ Z} is a natural constraint for Iε . Hence any critical point of Iε (z + wε (z)), z ∈ Z is a critical point of Iε . 3.3. Proof of Theorem 1 Throughout this section we will take ε and c is such a way that Proposition 1 applies. The main tool to prove Theorem 1 is the following Proposition Proposition 2. There results lim I c (zµ,ξ + wg (zµ,ξ )) = bc ,

µ→0

uniformly for ξ satisfying (5).

(23)

Hence I c |Zgc can be continuously extended to ∂Zgc by setting I c |∂Zgc = bc .

(24)

Postponing the proof of Proposition 2, it is immediate to deduce Theorem 1. Proof of Theorem 1. The extended functional I c has a critical point on the compact manifold Zgc ∪ ∂Zgc . From (24) it follows that either I c is identically constant or it achieves the maximum or the minimum in Zgc . In any case I c has a critical point on Zgc . According to Proposition 1, such a critical point gives rise to a solution of (Y ).   In order to prove Proposition 2 we prefer to reformulate (Y ) in a more convenient form using the stereographic projection σp , trough an appropriate point p ∈ ∂S+n , see Remark 3. In this way the problem reduces to study an elliptic equation in Rn+ , where calculation are easier. More precisely, let g˜ ij : Rn+ → R be the components of the metric g in σp -stereographic co-ordinates, and let

2 1 + |x|2 g ij = g˜ ij . (g) 2

680

A. Ambrosetti et al.

Then problem (Y ) is equivalent to find solutions of  (n − 1) n+2   −4 ∆g u + Rg u = u n−2 , in Rn+ ;    (n − 2)   2 n on ∂Rn+ = Rn−1 , ∂νg u + hg u = cu n−2 ,   (n − 2)     u > 0, u ∈ D1,2 (Rn ),

(Y )

+

where the symbols have obvious meaning. Solutions of problem (Y ) can be found as critical points of the functional fg : D1,2 (Rn+ ) → R defined in the following way    (n − 1) 1 1 ∗ 2 2 fg (u) = 2 |∇g u| dVg + Rg u dVg − ∗ u2 dVg n n n (n − 2) R+ 2 R+ 2 R+   n−1 + (n − 1) hg u2 dσg − c(n − 2) |u|2 n−2 dσg . ∂ Rn+

∂ Rn+

In general the transformation (g) induces an isometry between H 1 (B) and D1,2 (Rn+ ) given by

n−2 2 2 u(x) #→ u(x) := (x )2 + (xn + 1)2

2x (x )2 + xn2 − 1 × u , , (x )2 + (xn + 1)2 (x )2 + (xn + 1)2 where x = (x1 , . . . , xn−1 ). It turns out that fg (u) = I c (u) as well as

(25)

∇fg (u) = ∇I c (u).

In particular this implies that u solves (Y ) if and only if u is a solution of (Y ). Furthermore, there results – g0 corresponds to the trivial metric δij on Rn+ ; – z0 corresponds to z0 ∈ D1,2 (Rn+ ) given by z0 (x) = z0 (x − (0, a0 c)), c

x ∈ Rn+ ;

a0 =

κ ; 2

– Z c corresponds to Z given by 

 x − (ξ , a0 cµ) c − n−2 n−1 2 Z = zµ,ξ := µ z0 , µ > 0, ξ ∈ R . µ

Yamabe and scalar curvature

681 c

Let us point out that the manifold Z is nothing but τp ◦ τS−1 Z c (see Notations). From the preceding items it follows that the equation c

∇fg (z + w) ∈ Tz Z , c

have a unique solution w ⊥ Tz Z and there results w g (z) = wg (z). From this and (25) it follows I c (z + wg (z)) = fg (z + wg (z)).

(26)

Let us now introduce the metric g δ (x) := g(δx), δ > 0 and let fgδ : D1,2 (Rn+ ) → R be the corresponding Euler functional. For all u ∈ D1,2 (Rn+ ) there results

2−n fgδ (u) = fg δ 2 u(δ −1 x) . Introducing the linear isometry Tδ : D1,2 (Rn+ ) → D1,2 (Rn+ ) defined by Tδ (u) := n−2 δ − 2 u(x/δ) this becomes fgδ (u) = fg (Tδ u) ,

(27)

Furthermore, for all u ∈ D1,2 (Rn+ ) one has ∇fg (u) = Tδ ∇fgδ (Tδ−1 u) D fg (u)[v, w] = D 2

2

(28)

fgδ (Tδ−1 u)[Tδ−1 v, Tδ−1 w].

(29)

c

Arguing as above, there exists w gδ (z0 ) ∈ (Tz0 Z )⊥ such that c

∇fgδ (z0 + wgδ ) ∈ Tz0 Z . and there results wgδ (z0 )(x) = δ

n−2 2

wg (zδ )(δx),

namely wg (zδ ) = Tδ wgδ (z0 ).

(30)

Remark 2. From (27), (28), (29) and using the relations between fg and I c discussed above, it is easy to check that the estimates listed in Lemma 4 hold true, substituting I c with fgδ and z with z. A similar remark holds for Proposition 1.

682

A. Ambrosetti et al.

We are interested to the behaviour of fgδ as δ → 0. To this purpose, we set     1 ∗ 2 (n − 1) g ij (0)Di uDj u − ∗ |u|2  dVg(0) fg(0) (u) = (n − 2) i,j 2 Rn+  n−1 −c(n − 2) |u|2 n−2 dσg(0) , ∂ Rn+

which is the Euler functional corresponding to the constant metric g(0). Remark 3. Unlike the g δ , the metric g(0) does not come from a smooth metric on B. This is the main reason why it is easier to deal with (Y ) instead of (Y ). Lemma 5. For all u ∈ D1,2 (Rn+ ) there results lim ||∇fgδ (u) − ∇fg(0) (u)|| = 0;

(31)

lim fgδ (u) = fg(0) (u).

(32)

δ→0

δ→0

Proof. For any v ∈ D1,2 (Rn+ ) there holds   ∇fgδ (u) − ∇fg(0) (u), v = θ1 + θ2 + θ3 + θ4 + θ5 , where

   n−1 θ1 = 4 ∇gδ u · ∇gδ v dVgδ − ∇g(0) u · ∇g(0) v dVg(0) ; n−2 Rn+ Rn+  Rgδ u v dVgδ ; θ2 = Rn+   4 |u| n−2 u v (dVgδ − dVg(0) ); θ4 = 2(n − 1) hgδ u v dσgδ ; θ3 = Rn+

θ5 = 2c(n − 1)

 ∂ Rn+

|u|

2 n−2

∂ Rn−1

 u v dσgδ −

∂ Rn+

|u|

2 n−2



u v dσg(0) .

Using the Dominated Convergence Theorem and the integrability of |∇u|2 and ∗ of |u|2 , it is easy to show that θ1 , θ3 and θ5 converge to zero.  As far as θ2 is concerned, we first note that the bilinear form (u, v) → Rn Rg u v dVg is +

uniformly bounded for g ∈ G ε , so it turns out that given η > 0 there exists n uη ∈ Cc∞ (R+ ) such that        Rgδ u v dVgδ − Rgδ uη v dVgδ  ≤ η · v; ∀v ∈ D1,2 (Rn+ ). (33)  n  Rn+  R+

Yamabe and scalar curvature

683

Hence, since it is Rgδ (δ −1 x) = δ 2 Rg (x) (see (13)), it follows that for δ sufficiently small        2 R δ u v dVgδ  ≤ δ Rg L∞ (B) uη ∞ |v| = o(1) · v.   Rn+ g η  supp(uη ) So, using (33) and the arbitrarity of η, one deduces that θ2 = o(1) · v. Similar computations hold for the term θ4 . In the same way one can prove also (32).   We need a more complete description of w 0 (z). For this, according to Remark 3, we shall study the functional fg(0) in a direct fashion. If g ∈ Gε then the constant metric g(0) on Rn+ satisfies g(0) − I d∞ = O(ε) and thus fg(0) can be seen as a perturbation of the functional    (n − 1) 1 n−1 2 2∗ f0 (u) = 2 |∇u| dV0 − ∗ u dV0 − c(n − 2) |u|2 n−2 dσ0 , (n − 2) Rn+ 2 Rn+ ∂ Rn+ corresponding to the trivial metric δij . Then the procedure used in Sect. 3.2 yields to find w0 (z) such that c

(j) w 0 (z) is orthogonal to Tz Z ; c (jj) ∇fg(0) (z + w0 (z)) ∈ Tz Z ; n−2 c (jjj) w 0 (z) ≤ C · ε · (1 + |c|)− 2 ∀z ∈ Z . The following Lemma proves that a property stronger than (jj ) holds. c

Lemma 6. For all z ∈ Z there results ∇fg(0) (z + wg(0) (z)) = 0.

(34)

Hence z + wg(0) (z) solves 

(n−1) −4 (n−2) 2 ∂u (n−2) ∂ν

n

n+2

i,j =1

= cu

g ij (0)Dij2 u = u n−2

n n−2

in Rn+ ; on ∂Rn+ .

(35)

Here ν is the unit normal vector to ∂Rn+ with respect to g(0), namely g(0)(ν, ν) = 1;

g(0)(ν, v) = 0, ∀v ∈ ∂Rn+ .

Proof. The Lemma is a simple consequence of the invariance of the functional under the transformation Tµ,ξ : D1,2 (Rn+ ) → D1,2 (Rn+ ) defined in the following way

x − (ξ , 0) − n−2 2 Tµ,ξ (u) = µ . u µ This can be achieved with an elementary computation. It then follows that wg(0) (zµ,ξ ) = Tµ,ξ (wg(0) (z0 )),

for all µ, ξ .

684

A. Ambrosetti et al.

Hence, from the invariance of fg(0) , it turns out that fg(0) (zµ,ξ + w g(0) (zµ,ξ )) = fg(0) (Tµ,ξ (z0 + wg(0) (z0 ))) = fg(0) (z0 + wg(0) (z0 )). Since fg(0) (zµ,ξ + wg(0) (zµ,ξ )) is a constant function then, according to (j ) − (jj ), any z + wg(0) (z) is a critical point of fg(0) , proving the lemma.   Let us introduce some further notation: G denotes the matrix g ij (0), νg(0) is the outward unit normal to ∂Rn+ with respect to g ij (0), and e1 , . . . , en is the standard basis of Rn . Lemma 7. The solutions u of problem (35) are, up to dilations and translations, of the form u = z0 (Ax), where A is a matrix which satisfies −1

AG AT = I,

νg(0) =

 (A−1 )j n ej .

(36)

j

In particular, up to dilations, one has that z0 + w g(0) (z0 ) = z0 (A ·). Proof. First of all we prove the existence of a matrix A satisfying (36). The first −1 equality simply means that the bilinear form represented by the matrix G can be diagonalized, and this is standard. The matrix A which satisfies the first equation in (36) is defined uniquely up to multiplication on the left by an orthogonal matrix. Let (x1 , . . . , xn ) be the co-ordinates with respect to the standard basis (e1 , . . . , en ) of Rn , let (f1 , . . . , fn ) be the basis given by f = (A−1 )T e, and let (y1 , . . . , yn ) be the co-ordinates with respect to this new basis. This implies the relation between the co-ordinates x = Ay and the first of (36) implies that the bilinear form g ij (0) is diagonal with respect to y1 , . . . yn . Moreover, by the transitive action of O(n) over S n−1 we can ask that fn = ν; this is exactly the second equation in (36). In this way the matrix A is determined up to multiplication on the left by O(n − 1). We now prove that the function z˜ 0 = z0 (Ax) = z0 (y) is a solution of (35). First of all, since νg(0) is g(0)-orthogonal to ∂Rn−1 , the domain xn > 0 coincides with yn > 0 and the equation in the interior is, by formula (36) n n+2 (n − 1)  2 (n − 1)  ij −4 Dxi xj z˜ 0 (x) = −4 g Ali Akj Dy2k yl z0 (Ay) = z˜ 0n−2 (x). (n − 2) i,j =1 (n − 2) i,j

Yamabe and scalar curvature

685

Moreover, since ν = fn = ∂Rn+



j (A

−1 T )nj ej

=



j (A

−1

)j n ej , it turns out that on

 ∂ z˜ 0 (x) = (A−1 )j n Dxj z0 (Ay) ∂ν j  n (A−1 )j n Akj Dyk z0 (Ay) = Dyn z0 (Ay) = c˜z0n−2 (x). = j,k

Hence also the boundary condition is satisfied. Moreover, the function z0 ∈ D1,2 (Rn+ ) is the unique solution up to dilation and translation of problem (Y ) with g ij = I d, see [14]. As pointed out before, if A and A are two matrices satisfying (36), they differ up to O(n − 1). Then it is easy to check that z0 (Ax) = z0 (A x) and hence z˜ 0 is unique up to dilation and translation. This concludes the proof.   Corollary 1. The quantity fg(0) (z0 + w 0 (z0 )) is independent of g(0). Precisely one has: fg(0) (z0 + w 0 (z0 )) = bc . Proof. There holds fg(0) (z0 + wg(0) (z0 ))   (n − 1) g ij (0)Aki Alj Dk z0 (Ay)Dl z0 (Ay)dVg(0) (y) =2 (n − 2) Rn+ i,j,k,l   1 n−1 2∗ − ∗ |z0 (Ay)| dVg(0) (y) − c(n − 2) |z0 (Ay)|2 n−2 dσg(0) (y). n n 2 R+ ∂ R+ Using the change of variables x = Ay, and taking into account equations (12) and (36) we obtain the claim. This concludes the proof.   Lemma 8. There holds wgδ (z0 ) → wg(0) δ

as δ → 0.

(37)

c

Proof. Define H : D1,2 (Rn+ ) × Rn × Z → D1,2 (Rn+ ) × Rn by setting

δ ∇fgδ (z + w g(0) + w) − α z˙ H (w, α, z) = . ˙ (w, z) One has that ∇fgδ (z + wg(0) + w) = ∇fgδ (z + wg(0) ) + D 2 fgδ (z + wg(0) )[w] + ϑ(w) where



1

ϑ(w) := 0

 D 2 fgδ (z + wg(0) + sw) − D 2 fgδ (z + wg(0) ) [w]ds.



686

A. Ambrosetti et al. c

Recall that D 2 fgδ (z) is invertible on (Tz Z )⊥ . Since wg(0) satisfies (jjj ), then c also D 2 fgδ (z + wg(0) ) is invertible on (Tz Z )⊥ . As a consequence, the equation c c ∇fgδ (z + wg(0) + w) = 0, w ∈ (Tz Z )⊥ is equivalent, on (Tz Z )⊥ , to  −1   w = − D 2 fgδ (z + wg(0) ) ∇fgδ (z + wg(0) ) + ϑ(w) In addition, by Remark 2, we can use the estimates corresponding to (19) of Lemma 4 and to (iii) of Proposition 1, to infer that 

1

ϑ(w) = 0

 D 2 fgδ (z + wg(0) + sw) − D 2 fgδ (z + wg(0) ) [w]ds = o(w).



Then, repeating the arguments used in Sect. 3.2 with small changes, one can δ show that the equation H = 0 has a unique solution w = ω such that ω ≤ C · ∇fgδ (z + wg(0) ). From (34) and (31) it follows that ω → 0 as δ → 0. Since both wg(0) + ω c and wgδ solve (on (Tz Z )⊥ ) the same equation, we infer by uniqueness that wgδ = wg(0) + ω. Finally, since ω → 0 as δ → 0, then (37) follows.   Remark 4. All the preceding discussion has been carried out by taking the stereographic projection σp through an arbitrary p ∈ S n−1 . We are interested to the limit (23). When µ → 0 then ξ → ξ for some ξ ∈ S n−1 and it will be convenient to choose p = −ξ . We are now in position to give: Proof of Proposition 2. As pointed out in Remark 4, we take p = −ξ and use all the preceding results proved so far in this section. With this choice, when (µ, ξ ) → (0, ξ ) with ξ = |ξ | · ξ , zµ,ξ corresponds to zµ := zµ ,0 , for some µ → 0. Next, in view of (26), we will show that lim fg (zµ + wg (zµ )) = bc .

µ →0

By Corollary 1, bc = fg(0) (z0 + w g(0) ) and hence we need to prove that   lim fg (zµ + wg (zµ )) − fg(0) (z0 + w g(0) ) = 0. µ →0

Using (30), we have fg (zµ + wg (zµ )) = fg (zµ + Tµ w gµ (z0 )).

Yamabe and scalar curvature

687

Then we can write fg (zµ + w g (zµu )) − fg(0) (z0 + wg(0) ) = fg (zµ + Tµ w gµ (z0 )) = fg (zµ + Tµ wgµ (z0 )) − fg (zµ + Tµ wg(0) (z0 )) +ïfg (zµ + Tµ wg(0) (z0 )) − fg(0) (z0 + wg(0) ). From (17) with I c substituted by fg , we infer     fg (zµ + Tµ w gµ (z0 )) − fg (zµ + Tµ wg(0) ( ovz0 )) ≤ C · Tµ w gµ (z0 ) − Tµ wg(0) (z0 ) ≤ C · w gµ (z0 ) − wg(0) (z0 ) = o(1) as µ → 0.

Using zµ = Tµ z0 and (27), we deduce

fg (zµ + Tµ w gµ (z0 )) = fg Tµ (z0 + wgµ (z0 )) = fgµ (z0 + w g(0) ). Finally     fg (zµ + Tµ wgµ (z0 )) − fg(0) (z0 + w g(0) )     = fgµ (z0 + w g(0) ) − fg(0) (z0 + wg(0) ) → 0, according to Lemma 5. Since the above arguments can be carried out uniformly   with respect to ξ ∈ S n−1 , the proof is completed. 4. The scalar curvature problem In this section the value of c is fixed. Therefore its dependence will be omitted. So we will write Iε instead of Iεc , I0 instead of I0c , etc. 4.1. The abstract setting Solutions of problem (Pε ) can be found as critical points of the functional Iε : H 1 (B) → R defined as Iε (u) = I0 (u) − εG(u) where the unperturbed functional I0c (u) is defined by (see Sect. 2)   1 1 n−1 ∗ |u|2 − c(n − 2) |u|2 n−2 I0 (u) = u21 − ∗ 2 2 B S n−1

688

A. Ambrosetti et al.

and the perturbation G has the form   1 n−1 2∗ K(x)|u| dx + (n − 2) h(x)|u|2 n−2 dσ. G(u) = ∗ 2 B S n−1 The existence of critical points of Iε will be faced by means of the perturbation theory studied in [1]. Precisely, let us recall that I0 possesses an n-dimensional manifold Z = Z c , given by (6). Moreover, Z is non-degenerate in the sense that (i) − (ii) of Lemma 3 hold true. Then the results of [1] lead to consider the finite dimensional functional Γ := G|Z and give rise to the following Theorem: Theorem 5. In the preceding setting, let us suppose that either (a) Γ has a strict maximum (minimum) on Z; or (b) there exists an open subset Ω ⊂⊂ Z such that deg(Γ , Ω, 0) = 0. Then Iε has a critical point close to Z, provided ε is small enough. In our specific case, the function Γ (µ, ξ ) = G(zµ,ξ ) has the expression   (n−1) 2 (n−2) 1 2∗ Γ (µ, ξ ) = ∗ K(x)zµ,ξ (x)dx + (n − 2) h(σ )zµ,ξ (σ )dσ, 2 B S n−1

(38)

where µ > 0 and ξ ∈ Rn are related to c by (5), namely by µ2 + |ξ |2 − cκµ − 1 = 0. In order to apply the preceding abstract result we need to study the behaviour of Γ at the boundary of Z, which is given by ∂Z = {zµ,ξ0 : µ = 0, |ξ0 | = 1}. The following lemma will be proved in the Appendix and describes the behaviour of Γ at ∂Z. Below a1 , . . . , a6 denote positive constants defined in the Appendix. Lemma 9. Let |ξ0 | = 1 and let ν denote the outher normal direction to ∂Z at (0, ξ0 ). Γ can be extended to ∂Z and there results: (a) Γ (0, ξ0 ) = a1 K(ξ0 ) + a2 h(ξ0 ); (b) ∂ν Γ (0, ξ0 ) = a3 K (ξ0 ) · ξ0 ; (c) suppose that K (ξ0 ) · ξ0 = 0 and let n > 3. Then   ∂ν2 Γ (0, ξ0 ) = 4 a4 ∆T K(ξ0 ) + a5 D 2 K(ξ0 )[ξ0 , ξ0 ] + a6 ∆T h(ξ0 ) . Furthermore, if n = 3 and ∆T h(ξ0 ) = 0, then  +∞ provided ∆T h(ξ0 ) > 0, 2 ∂ν Γ (0, ξ0 ) = −∞ provided ∆T h(ξ0 ) < 0. The above Lemma is the counterpart of the calculation carried out in [2] for the Scalar Curvature Problem on S n .

Yamabe and scalar curvature

689

4.2. A general existence result Let us consider the auxiliary function ψ : S n−1 → R defined by ψ(x) = a1 K(x) + a2 h(x),

x ∈ S n−1 .

If x ∈ Crit (ψ) we denote by m(x, ψ) its Morse index. Theorem 6. Suppose that either (a) there exists an absolute maximum (resp. minimum) p ∈ S n−1 of ψ such that K (p) · p < 0 (resp. K (p) · p > 0); or (b) ψ is a Morse function satisfying K (x) · x = 0, 

∀ x ∈ Crit (ψ);

(39)

(−1)m(x,ψ) = 1.

(40)

x∈Crit (ψ), K (x)·x 0, there exists C > 0 such that for all c > −M there holds z ≤ C · (1 + |c|)−

n−2 2

for all z ∈ Z c .

(47)

Proof. By symmetry it suffices to take By symmetry it suffices to take ξ = 0 and consider z = zµ . As c → +∞ one has that µ ∼ κc and zµ ∼ µ(n−2)/2 in B. Then the lemma follows by a straight calculation.   Now we start by proving Eq. (15). Since it is clearly ∇I0c (z) = 0, it is sufficient to estimate the quantity ∇I c (z) − ∇I0c (z). Given v ∈ H 1 (B) and setting   (n − 1) (n − 1) ∇g z · ∇g v dVg − 4 ∇z · ∇v dV0 ; α1 = 4 (n − 2) B (n − 2) B  α2 = Rg z v dVg ; B   n+2 n+2 n−2 n−2 z v dV0 − z v dVg ; α4 = 2(n − 1) hg z v dσg ; α3 = B B ∂B   n n α5 = 2(n − 1) c z n−2 v dσg − 2(n − 1) c z n−2 v dσ0 , ∂B

there holds

∂B

 c  ∇I (z) − ∇I0c (z), v = α1 + α2 + α3 + α4 + α5 .

(48)

As far as α1 is concerned, taking into account of equations (12), (14) and the fact n−2 that z ≤ C · (1 + |c|)− 2 (see Lemma 10) one deduces that       ∇g z · ∇g v − ∇z · ∇v dx + C |∇z · ∇v| |dVg − dV0 | |α1 | ≤ C B

≤ C · ε · (1 + |c|)

− n−2 2

B

· v.

(49)

Turning to α2 we recall that the expression of Rg as a function of g, is of the type Rg = DΓ + G2 ;

Γ = Dg, ⇒ Rg = D 2 g + (Dg)2 .  We by estimating the quantity B Rg z v dV0 . Integrating by parts, the term  start 2 B D g z v dV0 transforms into    D 2 g z v dV0 = Dg z v dσ0 + Dg D(zv)dV0 . B

∂B

B

694

A. Ambrosetti et al.

Hence, if g ∈ Gε (see expression (2)), from the H¨older inequality it follows that   Rg zv dV0  (D 2 g + (Dg)2 )zv dV0 ≤ C · ε · z · v, B

B

and hence     n−2   Rg zv dV0 + |Rg zv||dVg − dV0 | ≤ C · ε · (1 + |c|)− 2 · v. |α2 | ≤ B

B

(50)

With simple estimates one can also prove that |α3 | ≤ C · ε · (1 + |c|)−

n+2 2

· v.

(51)

The function hg is of the form hg = Dg so, taking into account (2) one finds |α4 | ≤ C · ε · (1 + |c|)−

n−2 2

· v.

(52)

In order to estimate the last term α5 , using the continuous embedding H 1 (B) P→ n−1 L2 n−2 (S n−1 ) and the H¨older inequality one deduces that n

|α5 | ≤ C · ε · (1 + |c|) · z n−2n

L n−2 (S n−1 )

n

· v ≤ C · ε · (1 + |c|) · (1 + |c|)− 2 · v.

Putting together equations (49)-(52) one deduces (15). Turning to equation (19) and given v1 , v2 ∈ H 1 (B), there holds (D 2 I c (z + w) − D 2 I c (z))[v1 , v2 ] = δ1 + δ2 , where



 (n + 2) 4 4 u n−2 v1 v2 dVg − (u + w) n−2 v1 v2 dVg δ1 = (n − 2) B B 

 (n − 1) 2 2 n−2 n−2 c u v1 v2 dσg − (u + w) v1 v2 dσg . δ2 = 2n (n − 2) ∂B ∂B

Using standard inequalities one finds that  4 C · w n−2 for n ≥ 6,

|δ1 | ≤ 6−n 6−n C · w · u n−2 + w n−2 for n < 6;  4 C · (1 + |c|) · w n−2 for n ≥ 4,

|δ2 | ≤ 4−n 4−n C · (1 + |c|) · w · u n−2 + w n−2 for n < 4, D 2 I c (z + w) − D 2 I c (z) ≤ C · |c| · w n−2 . 2

(53)

Yamabe and scalar curvature

695

so we obtain the estimate. We now prove inequality (16). Given v1 , v2 ∈ H 1 (B) and setting   (n − 1) (n − 1) ∇g v1 · ∇g v2 dVg − 4 ∇v1 · ∇v2 dV0 ; β1 = 4 (n − 2) B (n − 2) B  β2 = Rg v1 v2 dVg ; B   (n + 2) (n + 2) 4 4 n−2 z v1 v2 dV0 − z n−2 v1 v2 dVg ; β3 = (n − 2) B (n − 2) B  hg v1 v2 dσg ; β4 = 2(n − 1) ∂B   (n − 1) (n − 1) 2 2 β5 = 2n c c z n−2 v1 v2 dσg − 2n z n−2 v1 v2 dσ0 , (n − 2) (n − 2) ∂B ∂B there holds (D 2 I c (z) − D 2 I0c (z))[v1 , v2 ] = β1 + β2 + β3 + β4 + β5 .

(54)

For β1 , taking into account equation (14) one finds         ∇g v1 · ∇g v2 − ∇v1 · ∇v2 dV0 + C |∇v1 · ∇v2 | · dVg − dV0  |β1 | ≤ C B

B

≤ C · ε · v1  · v2 .

(55)

Turning to β2 reasoning as for the above term α2 one deduces that    Rg z v  dVg ≤ C · ε · v1  · v2 . |β2 | ≤

(56)

B

In the same way one can prove that |β3 | ≤ C · ε · z n−2 · v1  · v2  ≤ C · ε · (1 + |c|)−2 · v1  · v2 . 4

(57)

For the term β4 , similarly to the expression α4 above there holds |β4 | ≤ C · ε · v1  · v2 .

(58)

Turning to β5 , using the H¨older inequality one deduces that 2

|β5 | ≤ C · c · ε · (1 + |c|) · z n−22

L n−2 (S n−1 )

· v1  · v2  ≤ C · ε · v1  · v2 . (59)

Putting together equations (55)-(59) (59) one deduces inequality (16). Equation (17) follows from similar computations.

696

A. Ambrosetti et al.

A.2. Proof of Lemma 9 Given ξ0 | = 1, we introduce a reference frame in Rn such that en = −ξ0 . Let α = α(µ) be such that the pair (µ, ξ ), with ξ = αξ0 , satisfies (5). Setting γ (µ) = Γ (µ, −α(µ)en ), one has that Γ (0, ξ0 ) = γ (0), ∂ν Γ (0, ξ0 ) = −γ (0), ∂ν2 Γ (0, ξ0 ) = γ (0). In order to evaluate the above quantities, it is convenient to make a change of variables. This will considerably simplify the calculation when we deal with γ (0) and γ (0). Let ψ : Rn+ → B be the map given by (y , yn ) ∈ Rn+ → (x , xn ) ∈ B; x =

2y (y )2 + yn2 − 1 , x = . n (y )2 + (yn + 1)2 (y )2 + (yn + 1)2

Here and in the sequel,if x ∈ Rn we will set x = (x1 , . . . , xn−1 ) so that x = (x , xn ). By using simple computations it turns out that γ (µ) = γ˜ (µ), ˜ where 1 γ˜ (µ) ˜ = ∗ 2

 Rn+

c 2∗ ˜ K(y)(z µ,0 ˜ ) (y)dy + (n − 2)

and µ˜ =

2µ ; 1+ + α(µ) µ2

 ∂ Rn+

n−1 c 2 n−2 ˜ h(ω)(z (ω)dω, µ,0 ˜ )

˜ K(y) = K(ψ(y)).

Let us point out that the derivatives of K and K˜ satisfy the following relations: ˜ 0) = 2Dxn K(ξ0 ); Dyn K(0, ˜ 0) = 2Dx (ξ0 ); Dy K(0,   2 ˜ Dyn K(0, 0) = 4 Dx2n K − Dxn K (ξ0 );   ˜ 0) = 4 Dx2 K − Dxn K (ξ0 ); Dy2 K(0,   ˜ 0) = 4 Dx2 ,xn K − Dx K (ξ0 ). Dy2 ,yn K(0, The change of variables y = µq, ˜ ω = µσ ˜ yields   1 n−1 c 2∗ c 2 n−2 ˜ ˜ µσ ) (σ )dσ. K(µq)(z ˜ h( ˜ )(z1,0 γ˜ (µ) ˜ = ∗ 1,0 ) (q)dq + (n − 2) 2 Rn+ ∂ Rn+ (60)

Yamabe and scalar curvature

697

Hence, passing to the limit for µ˜ → 0, it follows that ˜ ˜ + a2 h(0) = a1 K(ξ0 ) + a2 h(ξ0 ), γ (0) = γ˜ (0) = a1 K(0) with 1 a1 = ∗ 2

 Rn+



∗ z02 (q , qn

− κc/2)dq,

a2 = (n − 2)

2 n−1

∂ Rn+

z0 n−2 (σ, κc/2)dσ.

Let us now evaluate the first derivative. There holds d µ˜ d γ˜ (0) · (0) = 2γ˜ (0). d µ˜ dµ

γ (0) =

Moreover from formula (60) we deduce  1 ∗ c ˜ µq) ˜ = ∗ ∇ K( ˜ · q |z1,0 (q)|2 dq + (n − 2) γ˜ (µ) + 2  Rn n−1 c ˜ µσ × ∇ h( ˜ ) · σ |z1,0 (σ )|2 n−2 (σ )dσ. ∂ Rn+

(61)

For symmetry reasons when µ˜ → 0, the parallel component to ∂Rn+ in the first integral and the second integral vanishes, hence it follows that  2 ∗ c ˜ γ (0) = 2γ˜ (0) = ∗ Dn K(0) qn |z1,0 (q)|2 dq = −a3 K (ξ0 ) · ξ0 , (62) n 2 R+ where 4 a3 = ∗ 2





Rn+

qn z02 (q , qn − κc/2)dq.

We are interested in the study of the second derivative only in the case in which the first derivative vanishes, namely when K (ξ0 ) · ξ0 = 0. As for the second derivative, there holds: γ˜ (µ) ˜ =

1 2∗

 R+ n

n 



c 2 ˜ µq)q Dij2 K( ˜ i qj |z1,0 (q)| dq

i,j =1



+(n − 2)

n−1  ∂ Rn+ i,j =1

:= δ(µ) ˜ + ρ(µ). ˜

(n−1)

c ˜ µσ Dij2 h( (σ )|2 (n−2) dσ ˜ )σi σj |z1,0

(63)

Now we have to distinguish the case n = 3 and the case n > 3. In fact the boundary integral ρ(µ) ˜ in (63) is uniformly dominated by a function in L1 (∂Rn+ )

698

A. Ambrosetti et al.

if and only if n > 3. However it is possible to determine the sign of this integral also for n = 3: it turns out that  1 ∗ c ˜ lim δ(µ) ˜ = ∗ |q |2 |z1,0 (q)|2 dq · ∆T K(0) µ→0 ˜ 2 (n − 1) Rn+  1 ∗ c 2 ˜ qn2 |z1,0 (q)|2 dq · Dnn K(0); + ∗ 2 Rn+ and  ˜  ˜ = (+∞) · ∆T h(0), lim ρ(µu)  µ→0 ˜  (n−1) (n − 2)  c ˜  ˜ = |σ |2 |z1,0 (σ )|2 (n−2) dσ · ∆T h(0),  lim ρ(µ) µ→0 ˜ (n − 1) ∂ Rn+

for n = 3; for n > 3.

Hence we have that  for n = 3; (+∞) · ∆T h(ξ0 ) γ˜ (0) = (64)  2 a4 ∆T K(ξ0 ) + a5 D K(ξ0 )[ξ0 , ξ0 ] + a6 ∆T h(ξ0 ) for n > 3, where

 4 ∗ |q |2 z02 (q , qn − κc/2)dq, (n − 1)2∗ Rn+  4 ∗ qn2 z02 (q , qn − κc/2)dq, a5 = ∗ 2 Rn+  (n − 2) 2 n−1 |σ |2 z0 n−2 (σ, κc/2)dσ. a6 = 4 (n − 1) ∂ Rn+

a4 =

Finally, since γ (0) = 4γ˜ (0), the lemma follows. References 1. A. Ambrosetti, M. Badiale, Homoclinics: Poincar´e-Melnikov type results via a variational approach, Ann. Inst. Henri. Poincar´e Analyse Non Lin´eaire 15 (1998), 233–252. 323, (N +2)

2. A. Ambrosetti, J. Garcia Azorero, I. Peral, Perturbation of −∆u + u (N −2) = 0, the Scalar Curvature Problem in RN and related topics, J. Funct. Analysis, 165 (1999), 117–149. 3. A.Ambrosetti, A. Malchiodi, A multiplicity result for the Yamabe problem on S n , J. Funct. Analysis, 168 (1999), 529–561. 4. Ambrosetti A., Malchiodi A., On the Symmetric Scalar Curvature Problem on S n , J. Diff. Equations, 170-1 (2001), 228–245. 5. A.Ambrosetti, A. Malchiodi, Yan Yan Li, Scalar Curvature under Boundary Conditions, C. R. Acad. Sci. Paris, 330 (2000), 1013–1018. 6. T. Aubin, Some Nonlinear Problems in Differential Geometry, Springer, (1998).

Yamabe and scalar curvature

699

7. M.Berti, A. Malchiodi, Multiplicity results and multibump solutions for the Yamabe problem on S n , J. Funct. Anal. 180 (2001), 210–241. 8. A.Bahri, J. M. Coron, The Scalar-Curvature problem on the standard three-dimensional sphere, J. Funct. Anal. 95 (1991), 106–172. 9. A. Chang, X. Xu, P.Yang, A perturbation result for prescribing mean curvature, Math. Annalen 310 (1998), 473–496. 10. P. Cherrier, Problemes de Neumann nonlinearires sur les varietes Riemanniennes, J. Funct. Anal. 57 (1984), 154–207. 11. J. Escobar, Conformal deformation of a Riemannian metric to a scalar fla metric with constant mean curvature on the boundary, Ann. Math. 136 (1992), 1–50. 12. J. Escobar, Conformal metrics with prescribed mean curvature on the boundary, Cal. Var. 4 (1996), 559–592. 13. D.H. Gottlieb, A De Moivre like formula for fixed point theory, Contemp. Math. 72 (1988), 99–105. 14. Z.C.Han, Yan Yan Li, The Yamabe problem on manifolds with boundaries: existence and compactness results, Duke Math. J. 99 (1999), 489–542. 15. Z.C. Han, Yan Yan Li, The existence of conformal metrics with constant scalar curvature and constant boundary mean curvature, Comm. Anal. Geom., to appear. 16. Yan Yan Li, The Nirenberg problem in a domain with boundary, Top. Meth. Nonlin. Anal. 6 (1995), 309–329. 17. Yan Yan Li, Prescribing scalar curvature on S n and related topics, Part I, J. Diff. Equat. 120 (1995), 319–410; Part II, Existence and compactness, Comm. Pure Appl. Math. 49 (1996), 437–477. 18. Z. Djadli, A. Malchiodi, M.O. Ahmedou, Prescribing scalar and boundary mean curvature on the half three sphere, preprint. 19. M. Morse, G. Van Schaak, The critical point theory under general boundary conditions, Annals of Math. 35 (1934), 545–571.