Home On Gromov’s rigidity theorem for polytopes with acute angles
Article Open Access

On Gromov’s rigidity theorem for polytopes with acute angles

  • Simon Brendle and Yipeng Wang ORCID logo EMAIL logo
Published/Copyright: July 31, 2025

Abstract

In his “Four Lectures”, Gromov conjectured a scalar curvature extremality property of convex polytopes. Moreover, Gromov outlined a proof of the conjecture in the special case when the dihedral angles are acute. Gromov’s argument relies on Dirac operator techniques together with a smoothing construction. In this paper, we give the details of such a smoothing construction, thereby providing a detailed proof of Gromov’s theorem.

1 Introduction

In this paper, we prove a scalar curvature rigidity property for polytopes with acute angles. Suppose that n 3 is an integer, and Ω is a compact, convex polytope in R n with non-empty interior. We write Ω = m = 0 q { u m 0 } , where u 0 , , u q are non-constant linear functions defined on R n . After eliminating redundant inequalities, we may assume that the following condition is satisfied.

Assumption 1.1

For each 0 k q , the set

{ u k > 0 } m { 0 , 1 , , q } { k } { u m 0 }

is non-empty.

Without loss of generality, we assume that the following condition is satisfied.

Assumption 1.2

For each 0 k q , the gradient of the function u k with respect to the Euclidean metric is given by a unit vector N k S n 1 .

Theorem 1.3

Theorem 1.3 (cf. M. Gromov [4, Section 3.18])

Suppose that n 3 is an integer, and

Ω = m = 0 q { u m 0 }

is a compact, convex polytope in R n with non-empty interior. Suppose that Assumptions 1.1 and 1.2 are satisfied. Let 𝑔 be a Riemannian metric on R n . For each k { 0 , 1 , , q } , we denote by ν k = u k / | u k | the unit normal vector to the level sets of u k with respect to the metric 𝑔. We assume that the following conditions are satisfied.

  • The scalar curvature of 𝑔 is nonnegative at each point in Ω.

  • For each k { 0 , 1 , , q } , the mean curvature of the hypersurface { u k = 0 } with respect to 𝑔 is nonnegative at each point in Ω { u k = 0 } .

  • If 𝑗 and 𝑘 are integers with 0 j < k q and 𝑥 is a point in Ω with u j ( x ) = u k ( x ) = 0 , then ν j , ν k N j , N k 0 at the point 𝑥.

Then
  • the Riemann curvature tensor of 𝑔 vanishes at each point in Ω.

  • For each k { 0 , 1 , , q } , the second fundamental form of the hypersurface { u k = 0 } with respect to 𝑔 vanishes at each point in Ω { u k = 0 } .

  • If 𝑗 and 𝑘 are integers with 0 j < k q and 𝑥 is a point in Ω with u j ( x ) = u k ( x ) = 0 , then ν j , ν k = N j , N k at the point 𝑥.

Theorem 1.3 is a special case of Gromov’s Dihedral Rigidity Conjecture. Theorem 1.3 was stated by Gromov in his “Four Lectures” (see [4, Section 3.18]). Moreover, Gromov sketched a proof based on Dirac operator techniques and a smoothing construction. The main purpose of this paper is to work out the details of the smoothing construction.

In Section 2, we explain how a given polytope Ω may be approximated by a convex domain Ω ̂ Ω with smooth boundary Ω ̂ = Σ ̂ .

In Section 3, we construct a smooth map N ̂ : Σ ̂ S n 1 which is homotopic to the Gauss map of Σ ̂ .

In Section 4, we explain how to control certain angles under the smoothing procedure (see Lemma 4.3 and Proposition 4.4 below). Here, we use in an essential way our assumption that Ω has acute angles.

In Section 5, we prove pointwise estimates for max { d N ̂ tr H , 0 } in various subsets of Σ ̂ , where 𝐻 denotes the mean curvature of Σ ̂ with respect to the metric 𝑔. From this, we deduce that a suitable Morrey norm of max { d N ̂ tr H , 0 } can be made arbitrarily small (see Corollary 5.10 below).

In Section 6, we complete the proof of Theorem 1.3. The proof follows the arguments in [1]. It uses the bound for the Morrey norm of max { d N ̂ tr H , 0 } established in Section 5, together with a deep estimate due to Fefferman and Phong [2].

We refer to [1, 3, 4, 5, 6, 7] for related work on Gromov’s Dihedral Rigidity Conjecture. We note that Wang, Xie, and Yu [8] have proposed a different approach to this problem which is based on the study of Dirac operators on manifolds with corners; their work is currently in the process of verification.

2 Smoothing convex polytopes with acute angles

Throughout this section, we assume that n 3 is an integer, and Ω is a compact, convex polytope in R n with non-empty interior. As above, we write Ω = m = 0 q { u m 0 } , where u 0 , , u q are non-constant linear functions defined on R n . Throughout this section, we assume that Assumptions 1.1 and 1.2 are satisfied. Moreover, we assume that the following assumption is satisfied.

Assumption 2.1

Let 0 j < k q . If there exists a point x Ω satisfying

u j ( x ) = u k ( x ) = 0 ,

then N j , N k 0 .

Lemma 2.2

Given ε ( 0 , 1 ) , we can find a small positive real number 𝛿 (depending on 𝜀) with the following property. Suppose that 0 j < k q . If there exists a point x Ω with δ u j ( x ) 0 and δ u k ( x ) 0 , then N j , N k ε .

Proof

Suppose that the assertion is false. We consider a sequence of counterexamples and pass to the limit. In the limit, we obtain a pair of integers 0 j < k q and a point x Ω such that u j ( x ) = u k ( x ) = 0 and N j , N k ε . This contradicts Assumption 2.1. ∎

Lemma 2.3

We can find a large constant Λ > 1 with the following property. Suppose that 𝑥 is a point in Ω and a 0 , , a q are nonnegative real numbers such that a i = 0 for all 0 i q satisfying u i ( x ) 2 Λ 1 . Then i = 0 q a i Λ | i = 0 q a i N i | .

Proof

Suppose that the assertion is false. We consider a sequence of counterexamples and pass to the limit. In the limit, we obtain a point x Ω and a collection of nonnegative real numbers a 0 , , a q with the following properties.

  • i = 0 q a i = 1 .

  • i = 0 q a i N i = 0 .

  • a i = 0 for all 0 i q satisfying u i ( x ) < 0 .

For abbreviation, let I : = { 0 , 1 , , q } and I 0 : = { i I : u i ( x ) = 0 } . Since Ω is a convex set with non-empty interior, we can find a vector ξ R n such that N i , ξ > 0 for all i I 0 . On the other hand, since a i = 0 for all i I I 0 , we obtain i I 0 a i N i , ξ = i I a i N i , ξ = 0 and i I 0 a i = i I a i = 1 . This is a contradiction. ∎

Lemma 2.4

We can find a large constant Ξ ( 2 Λ , ) with the following property. Suppose that 1 k q , and suppose that 𝑥 is a point in Ω with

2 Ξ 1 u k ( x ) 0 .

Moreover, suppose that a 0 , , a k 1 are nonnegative real numbers such that a i = 0 for all 0 i k 1 satisfying u i ( x ) 2 Ξ 1 . Then

i = 0 k 1 a i Ξ | ( i = 0 k 1 a i N i ) N k | .

Proof

Let Λ denote the constant in Lemma 2.3. Let us fix a positive real number δ ( 0 , Λ 1 ) so that the conclusion of Lemma 2.2 holds with ε = 1 2 Λ 1 . Let us define

Ξ = 2 δ 1 ( 2 Λ , ) .

We claim that Ξ has the desired property. To prove this, suppose that 1 k q , and suppose that 𝑥 is a point in Ω with δ u k ( x ) 0 . Moreover, suppose that a 0 , , a k 1 are nonnegative real numbers such that a i = 0 for all 0 i k 1 satisfying u i ( x ) δ . We define

b = i = 0 k 1 a i N i , N k .

Clearly,

| ( i = 0 k 1 a i N i ) N k | = | i = 0 k 1 a i N i + b N k | .

To estimate the term on the right-hand side, we distinguish two cases.

Case 1.  Suppose first that b 0 . Lemma 2.3 implies

| i = 0 k 1 a i N i + b N k | Λ 1 ( i = 0 k 1 a i + b ) Λ 1 i = 0 k 1 a i .

Case 2.  Suppose next that b 0 . Using Lemma 2.3, we obtain

| i = 0 k 1 a i N i | Λ 1 i = 0 k 1 a i .

Moreover, it follows from Lemma 2.2 that N i , N k 1 2 Λ 1 for all 0 i k 1 satisfying δ u i ( x ) 0 . This implies

b = i = 0 k 1 a i N i , N k 1 2 Λ 1 i = 0 k 1 a i .

Consequently,

| i = 0 k 1 a i N i + b N k | | i = 0 k 1 a i N i | + b 1 2 Λ 1 i = 0 k 1 a i .

This completes the proof of Lemma 2.4. ∎

Throughout this paper, we fix a smooth, even function η : R R such that η ( t ) = | t | for | t | 1 2 and η ′′ ( t ) 0 for all t R . Note that 1 η ( t ) 0 for t 0 and 0 η ( t ) 1 for t 0 . Moreover, | t | η ( t ) | t | + 1 for all t R .

In the following, γ ( 0 , 1 2 ) will denote a small parameter, and λ 0 > 1 will denote a large parameter. For abbreviation, we put λ k : = γ k λ 0 for 1 k q .

Definition 2.5

We define a collection of smooth functions u ̂ 0 , , u ̂ q on R n so that u ̂ 0 = u 0 and

u ̂ k = 1 2 ( u ̂ k 1 + u k + λ k 1 η ( λ k ( u ̂ k 1 u k ) ) )

for 1 k q . Moreover, we define Ω ̂ : = { u ̂ q 0 } and Σ ̂ : = { u ̂ q = 0 } .

Lemma 2.6

We have

max { u ̂ k 1 , u k } u ̂ k max { u ̂ k 1 , u k } + λ k 1

for 1 k q . In particular, u ̂ 0 u ̂ 1 u ̂ q .

Proof

This follows from the fact that | t | η ( t ) | t | + 1 for all t R . ∎

Lemma 2.7

We have

max { u 0 , , u q } u ̂ q max { u 0 , , u q } + k = 1 q λ k 1 .

In particular, m = 0 q { u m λ 0 1 } Ω ̂ Ω .

Proof

This follows immediately from Lemma 2.6. ∎

Lemma 2.8

Let 0 k q . Then | d u ̂ k | C at each point in R n . Moreover, u ̂ k is a convex function and the Hessian of u ̂ k is bounded by C λ k at each point in R n . Here, 𝐶 is independent of 𝛾 and λ 0 .

Proof

The proof is by induction on 𝑘. The assertion is obvious for k = 0 . Suppose next that 1 k q and that the assertion is true for u ̂ k 1 . The differential of u k is given by

d u ̂ k = 1 2 ( 1 + η ( λ k ( u ̂ k 1 u k ) ) ) d u ̂ k 1 + 1 2 ( 1 η ( λ k ( u ̂ k 1 u k ) ) ) d u k .

Moreover,

D 2 u ̂ k = 1 2 ( 1 + η ( λ k ( u ̂ k 1 u k ) ) ) D 2 u ̂ k 1 + 1 2 λ k η ′′ ( λ k ( u ̂ k 1 u k ) ) ( d u ̂ k 1 d u k ) ( d u ̂ k 1 d u k ) ,

where D 2 u ̂ k denotes the Hessian of u ̂ k with respect to the Euclidean metric and D 2 u ̂ k 1 denotes the Hessian of u ̂ k 1 with respect to the Euclidean metric. From this, we deduce that the assertion is true for u ̂ k . This completes the proof of Lemma 2.8. ∎

Lemma 2.9

Let 0 k q and let 𝑥 be a point in Ω. Then we can find nonnegative real numbers a 0 , , a k such that i = 0 k a i = 1 and d u ̂ k = i = 0 k a i d u i at the point 𝑥. Moreover, a i = 0 for all 0 i k satisfying u i ( x ) < u ̂ k ( x ) 2 k λ i 1 .

Proof

We argue by induction on 𝑘. The assertion is clearly true for k = 0 . We next assume that 1 k q , and the assertion is true for k 1 . To prove the assertion for 𝑘, we fix an arbitrary point x Ω .

Case 1.  Suppose that u ̂ k 1 ( x ) u k ( x ) > λ k 1 . In this case, u ̂ k = u ̂ k 1 in an open neighborhood of 𝑥. In view of the induction hypothesis, we can find nonnegative real numbers b 0 , , b k 1 such that i = 0 k 1 b i = 1 and d u ̂ k 1 = i = 0 k 1 b i d u i at the point 𝑥. Moreover, b i = 0 for all 0 i k 1 satisfying u i ( x ) < u ̂ k 1 ( x ) 2 ( k 1 ) λ i 1 . We define nonnegative real numbers a 0 , , a k by a i = b i for 0 i k 1 and a k = 0 . Then

i = 0 k a i = i = 0 k 1 b i = 1 and d u ̂ k = d u ̂ k 1 = i = 0 k 1 b i d u i = i = 0 k a i d u i

at the point 𝑥. Moreover, a i = 0 for all 0 i k satisfying u i ( x ) < u ̂ k ( x ) 2 k λ i 1 .

Case 2.  Suppose that u ̂ k 1 ( x ) u k ( x ) < λ k 1 . In this case, u ̂ k = u k in an open neighborhood of 𝑥. We define nonnegative real numbers a 0 , , a k by a i = 0 for 0 i k 1 and a k = 1 . Then i = 0 k a i = 1 and d u ̂ k = d u k = i = 0 k a i d u i . Moreover, a i = 0 for all 0 i k satisfying u i ( x ) < u ̂ k ( x ) 2 k λ i 1 .

Case 3.  Suppose that | u ̂ k 1 ( x ) u k ( x ) | λ k 1 . In this case, Lemma 2.6 implies

u ̂ k ( x ) max { u ̂ k 1 ( x ) , u k ( x ) } + λ k 1 min { u ̂ k 1 ( x ) , u k ( x ) } + 2 λ k 1 .

In view of the induction hypothesis, we can find nonnegative real numbers b 0 , , b k 1 such that i = 0 k 1 b i = 1 and d u ̂ k 1 = i = 0 k 1 b i d u i at the point 𝑥. Moreover, we have b i = 0 for all 0 i k 1 satisfying u i ( x ) < u ̂ k 1 ( x ) 2 ( k 1 ) λ i 1 . We define nonnegative real numbers a 0 , , a k by

a i = 1 2 ( 1 + η ( λ k ( u ̂ k 1 u k ) ) ) b i

for 0 i k 1 , and

a k = 1 2 ( 1 η ( λ k ( u ̂ k 1 u k ) ) ) .

Then

i = 0 k a i = 1 2 ( 1 + η ( λ k ( u ̂ k 1 u k ) ) ) i = 0 k 1 b i + 1 2 ( 1 η ( λ k ( u ̂ k 1 u k ) ) ) = 1 2 ( 1 + η ( λ k ( u ̂ k 1 u k ) ) ) + 1 2 ( 1 η ( λ k ( u ̂ k 1 u k ) ) ) = 1 ,
d u ̂ k = 1 2 ( 1 + η ( λ k ( u ̂ k 1 u k ) ) ) d u ̂ k 1 + 1 2 ( 1 η ( λ k ( u ̂ k 1 u k ) ) ) d u k = 1 2 ( 1 + η ( λ k ( u ̂ k 1 u k ) ) ) i = 0 k 1 b i d u i + 1 2 ( 1 η ( λ k ( u ̂ k 1 u k ) ) ) d u k = i = 0 k a i d u i
at the point 𝑥. Finally, a i = 0 for all 0 i k satisfying u i ( x ) < u ̂ k ( x ) 2 k λ i 1 . This completes the proof of Lemma 2.9. ∎

Lemma 2.10

If λ 0 is sufficiently large, then the following holds. Let 0 k q . If 𝑥 is a point in Ω satisfying Λ 1 u ̂ k ( x ) 0 , then | d u ̂ k | Λ 1 at the point 𝑥. In particular, Σ ̂ is a smooth hypersurface.

Proof

This follows by combining Lemma 2.3 and Lemma 2.9. ∎

Lemma 2.11

If λ 0 is sufficiently large, then the following holds. Let 0 j < k q . If 𝑥 is a point in Ω satisfying Ξ 1 u ̂ j ( x ) 0 , Ξ 1 u k ( x ) 0 , then | d u ̂ j d u k | Ξ 1 at the point 𝑥.

Proof

Consider a point x Ω satisfying Ξ 1 u ̂ j ( x ) 0 and Ξ 1 u k ( x ) 0 . By Lemma 2.9, we can find nonnegative real numbers a 0 , , a j such that i = 0 j a i = 1 and d u ̂ j = i = 0 j a i d u i at the point 𝑥. Moreover, a i = 0 for all i { 0 , 1 , , j } satisfying u i ( x ) < u ̂ j ( x ) 2 j λ i 1 . As Ξ 1 u ̂ j ( x ) 0 , it follows that a i = 0 for all i { 0 , 1 , , j } satisfying u i ( x ) 2 Ξ 1 . Using Assumption 1.2 and Lemma 2.4, we obtain

1 = i = 0 j a i Ξ | ( i = 0 j a i d u i ) d u k | = Ξ | d u ̂ j d u k |

at the point 𝑥. This completes the proof of Lemma 2.11. ∎

Lemma 2.12

Let 1 k q . Then

| Ω { u ̂ k = 0 } { 2 λ k 1 u ̂ k 1 0 } { 2 λ k 1 u k 0 } B r ( p ) | C λ k 1 r n 2

for all 0 < r 1 . Here, the expression on the left-hand side represents the ( n 1 ) -dimensional measure. Moreover, B r ( p ) denotes a Euclidean ball of radius 𝑟, and 𝐶 is independent of 𝛾 and λ 0 .

Proof

Lemma 2.11 implies that | d u ̂ k 1 d u k | Ξ 1 at each point in

Ω { 2 λ k 1 u ̂ k 1 0 } { 2 λ k 1 u k 0 } .

Let us consider two arbitrary real numbers a , b [ 2 λ k 1 , 0 ] . We apply Proposition A.1 with m = n 1 and with 𝑓 defined as the restriction of the function u ̂ k 1 a to the hyperplane { u k = b } . This gives

| { d u ̂ k 1 d u k 0 } { u ̂ k 1 = a } { u k = b } B r ( p ) | C ( n ) r n 2

for all 0 < r 1 , where the expression on the left-hand side represents the ( n 2 ) -dimensional measure. Since d u ̂ k 1 d u k 0 at each point in

Ω { 2 λ k 1 u ̂ k 1 0 } { 2 λ k 1 u k 0 } ,

it follows that

(2.1) | Ω { u ̂ k 1 = a } { u k = b } B r ( p ) | C ( n ) r n 2

for all a , b [ 2 λ k 1 , 0 ] and all 0 < r 1 , where the expression on the left-hand side represents the ( n 2 ) -dimensional measure.

For abbreviation, let

S k = Ω { u ̂ k = 0 } { 2 λ k 1 u ̂ k 1 0 } { 2 λ k 1 u k 0 } .

Moreover, we put S k + = S k { u ̂ k 1 u k 0 } and S k = S k { u ̂ k 1 u k 0 } .

For each b [ 2 λ k 1 , 0 ] , we can find a real number a [ b , 0 ] (depending on 𝑏) such that S k + { u k = b } { u ̂ k 1 = a } . Using (2.1), we obtain

(2.2) | S k + { u k = b } B r ( p ) | C ( n ) r n 2

for each b [ 2 λ k 1 , 0 ] and all 0 < r 1 , where the expression on the left-hand side represents the ( n 2 ) -dimensional measure. Similarly, for each a [ 2 λ k 1 , 0 ] , we can find a real number b [ a , 0 ] (depending on 𝑎) such that S k { u ̂ k 1 = a } { u k = b } . Using (2.1), we obtain

(2.3) | S k { u ̂ k 1 = a } B r ( p ) | C ( n ) r n 2

for each a [ 2 λ k 1 , 0 ] and all 0 < r 1 , where the expression on the left-hand side represents the ( n 2 ) -dimensional measure. In the next step, we integrate inequality (2.2) over b [ 2 λ k 1 , 0 ] , and we integrate inequality (2.3) over a [ 2 λ k 1 , 0 ] . Using the co-area formula, we obtain

S k + B r ( p ) | d u k d u ̂ k | | d u ̂ k | C ( n ) λ k 1 r n 2 , S k B r ( p ) | d u ̂ k 1 d u ̂ k | | d u ̂ k | C ( n ) λ k 1 r n 2

for all 0 < r 1 . It follows from Lemma 2.8 that | d u ̂ k | C at each point in Ω. Using the identity

d u ̂ k = 1 2 ( 1 + η ( λ k ( u ̂ k 1 u k ) ) ) d u ̂ k 1 + 1 2 ( 1 η ( λ k ( u ̂ k 1 u k ) ) ) d u k ,

we obtain

| d u k d u ̂ k | = 1 2 ( 1 + η ( λ k ( u ̂ k 1 u k ) ) ) | d u k d u ̂ k 1 | 1 2 | d u k d u ̂ k 1 | 1 2 Ξ 1

at each point in S k + and

| d u ̂ k 1 d u ̂ k | = 1 2 ( 1 η ( λ k ( u ̂ k 1 u k ) ) ) | d u ̂ k 1 d u k | 1 2 | d u ̂ k 1 d u k | 1 2 Ξ 1

at each point in S k . Putting these facts together, we conclude that

| S k + B r ( p ) | C λ k 1 r n 2 and | S k B r ( p ) | C λ k 1 r n 2

for all 0 < r 1 . Thus, | S k B r ( p ) | C λ k 1 r n 2 for all 0 < r 1 . This completes the proof of Lemma 2.12. ∎

Lemma 2.13

Suppose that i , j , k { 0 , 1 , , q } are pairwise distinct. If λ 0 is sufficiently large, then

| Ω { u ̂ k = 0 } { 2 λ k 1 u ̂ k 1 0 } { 2 λ k 1 u k 0 } { 4 λ 0 1 u j 0 } { 6 λ 0 1 u i 0 } B r ( p ) | C λ k 1 λ 0 1 r n 3

for all λ 0 1 r 1 . Here, the expression on the left-hand side represents the ( n 1 ) -dimensional measure. Moreover, B r ( p ) denotes a Euclidean ball of radius 𝑟, and 𝐶 is independent of 𝛾 and λ 0 .

Proof

As above, we define

S k = Ω { u ̂ k = 0 } { 2 λ k 1 u ̂ k 1 0 } { 2 λ k 1 u k 0 } .

We distinguish two cases.

Case 1.  Suppose that Ω { u i = 0 } { u j = 0 } { u k = 0 } = . Then

Ω { 2 λ 0 1 u k 0 } { 4 λ 0 1 u j 0 } { 6 λ 0 1 u i 0 } =

if λ 0 is sufficiently large. Consequently,

S k { 4 λ 0 1 u j 0 } { 6 λ 0 1 u i 0 } =

if λ 0 is sufficiently large. Thus, the assertion is trivially true in this case.

Case 2.  Suppose that Ω { u i = 0 } { u j = 0 } { u k = 0 } . It follows from Assumption 1.1 that the hyperplanes { u i = 0 } , { u j = 0 } , { u k = 0 } must intersect transversally. Suppose now that λ 0 1 r 1 . We can cover the set

{ 2 λ 0 1 u k 0 } { 4 λ 0 1 u j 0 } { 6 λ 0 1 u i 0 } B r ( p )

by C ( λ 0 r ) n 3 balls of radius λ 0 1 . By Lemma 2.12, the intersection of S k with each such ball has area at most C λ k 1 λ 0 2 n . Therefore, the set

S k { 4 λ 0 1 u j 0 } { 6 λ 0 1 u i 0 } B r ( p )

has area at most C λ k 1 λ 0 1 r n 3 . This completes the proof of Lemma 2.13. ∎

Definition 2.14

We define

F 0 = Σ ̂ m = 1 q { u ̂ m 1 u m > λ m 1 } .

For 1 k q , we define

F k = Σ ̂ m = k + 1 q { u ̂ m 1 u m > λ m 1 } { u ̂ k 1 u k < λ k 1 } .

Definition 2.15

For 1 k q , we define

E 0 , k = Σ ̂ m = k + 1 q { u ̂ m 1 u m > λ m 1 } m = 1 k 1 { u ̂ m 1 u m > λ m 1 } { 2 λ k 1 u ̂ k 1 0 } { 2 λ k 1 u k 0 } { 2 λ k 1 u 0 0 } .

For 1 j < k q , we define

E j , k = Σ ̂ m = k + 1 q { u ̂ m 1 u m > λ m 1 } m = j + 1 k 1 { u ̂ m 1 u m > λ m 1 } { 2 λ k 1 u ̂ k 1 0 } { 2 λ k 1 u k 0 } { 2 λ k 1 u j 0 } { u ̂ j 1 u j < λ j 1 } .

Definition 2.16

For 0 i < j < k q , we define

G i , j , k = Σ ̂ m = k + 1 q { u ̂ m 1 u m > λ m 1 } m = j + 1 k 1 { u ̂ m 1 u m > λ m 1 } { 2 λ k 1 u ̂ k 1 0 } { 2 λ k 1 u k 0 } { 4 λ j 1 u j 0 } { 6 λ i 1 u i 0 } .

Proposition 2.17

We have

Σ ̂ = 0 k q F k 0 j < k q E j , k 0 i < j < k q G i , j , k .

Proof

Let us consider an arbitrary point x Σ ̂ . We claim that

x 0 k q F k 0 j < k q E j , k 0 i < j < k q G i , j , k .

If u ̂ m 1 ( x ) u m ( x ) > λ m 1 for all m { 1 , , q } , then x F 0 . Otherwise, we define an integer k { 1 , , q } by

k = max { m { 1 , , q } : u ̂ m 1 ( x ) u m ( x ) λ m 1 } .

Clearly, u ̂ k 1 ( x ) u k ( x ) λ k 1 and u ̂ m 1 ( x ) u m ( x ) > λ m 1 for all m { k + 1 , , q } . From this, we deduce that u ̂ k ( x ) = u ̂ q ( x ) = 0 . If u ̂ k 1 ( x ) u k ( x ) < λ k 1 , then x F k .

It remains to consider the case when | u ̂ k 1 ( x ) u k ( x ) | λ k 1 . Since u ̂ k ( x ) = 0 , Lemma 2.6 implies that max { u ̂ k 1 ( x ) , u k ( x ) } [ λ k 1 , 0 ] . Therefore, u ̂ k 1 ( x ) [ 2 λ k 1 , 0 ] and u k ( x ) [ 2 λ k 1 , 0 ] . If u ̂ m 1 ( x ) u m ( x ) > λ m 1 for all m { 1 , , k 1 } , then

u 0 ( x ) = u ̂ 0 ( x ) = u ̂ k 1 ( x ) [ 2 λ k 1 , 0 ] ,

and consequently x E 0 , k . Otherwise, we define an integer j { 1 , , k 1 } by

j = max { m { 1 , , k 1 } : u ̂ m 1 ( x ) u m ( x ) λ m 1 } .

Clearly, u ̂ j 1 ( x ) u j ( x ) λ j 1 and u ̂ m 1 ( x ) u m ( x ) > λ m 1 for all m { j + 1 , , k 1 } . From this, we deduce that u ̂ j ( x ) = u ̂ k 1 ( x ) [ 2 λ k 1 , 0 ] . If u ̂ j 1 ( x ) u j ( x ) < λ j 1 , then u j ( x ) = u ̂ j ( x ) = u ̂ k 1 ( x ) [ 2 λ k 1 , 0 ] , and consequently x E j , k .

It remains to consider the case when | u ̂ j 1 ( x ) u j ( x ) | λ j 1 . As u ̂ j ( x ) [ 2 λ k 1 , 0 ] , Lemma 2.6 implies that max { u ̂ j 1 ( x ) , u j ( x ) } [ 3 λ j 1 , 0 ] . Therefore, u ̂ j 1 ( x ) [ 4 λ j 1 , 0 ] and u j ( x ) [ 4 λ j 1 , 0 ] . If u ̂ m 1 ( x ) u m ( x ) > λ m 1 for all m { 1 , , j 1 } , then

u 0 ( x ) = u ̂ 0 ( x ) = u ̂ j 1 ( x ) [ 4 λ j 1 , 0 ] ,

and consequently x G 0 , j , k . Otherwise, we define an integer i { 1 , , j 1 } by

i = max { m { 1 , , j 1 } : u ̂ m 1 ( x ) u m ( x ) λ m 1 } .

Clearly, u ̂ i 1 ( x ) u i ( x ) λ i 1 and u ̂ m 1 ( x ) u m ( x ) > λ m 1 for all m { i + 1 , , j 1 } . From this, we deduce that u ̂ i ( x ) = u ̂ j 1 ( x ) [ 4 λ j 1 , 0 ] . Using Lemma 2.6, we obtain

max { u ̂ i 1 ( x ) , u i ( x ) } [ 5 λ i 1 , 0 ] .

Since u ̂ i 1 ( x ) u i ( x ) λ i 1 , we conclude that u i ( x ) [ 6 λ i 1 , 0 ] , and consequently we get x G i , j , k . This completes the proof of Proposition 2.17. ∎

3 The map N ̂ : Σ ̂ S n 1

Suppose that n 3 is an integer, and Ω = m = 0 q { u m 0 } is a compact, convex polytope in R n with non-empty interior. Throughout this section, we assume that Assumptions 1.1, 1.2, and 2.1 are satisfied. Moreover, we assume that 𝑔 is an arbitrary Riemannian metric on R n .

Definition 3.1

Let 0 k q . At each point x R n , we define ν k = u k / | u k | , where u k and | u k | are computed with respect to the metric 𝑔. At each point

x R n { d u ̂ k = 0 } ,

we define ν ̂ k = u ̂ k / | u ̂ k | , where u ̂ k and | u ̂ k | are computed with respect to the metric 𝑔. Note that Ω { 2 k Ξ 1 u ̂ k 0 } R n { d u ̂ k = 0 } by Lemma 2.10. Finally, we denote by ν ̂ the restriction of the vector field ν ̂ q to the hypersurface Σ ̂ = { u ̂ q = 0 } .

Proposition 3.2

If λ 0 is sufficiently large, then the following statements hold.

  1. Let 1 k q . Suppose that 𝑥 is a point in Ω satisfying 2 k Ξ 1 u ̂ k ( x ) 0 and u ̂ k 1 ( x ) u k ( x ) > λ k 1 . Then 2 k + 1 Ξ 1 u ̂ k 1 ( x ) 0 . Moreover, ν ̂ k = ν ̂ k 1 at the point 𝑥.

  2. Let 1 k q . Suppose that 𝑥 is a point in Ω satisfying 2 k Ξ 1 u ̂ k ( x ) 0 and u ̂ k 1 ( x ) u k ( x ) < λ k 1 . Then ν ̂ k = ν k at the point 𝑥.

  3. Let 1 k q . Suppose that 𝑥 is a point in Ω satisfying 2 k Ξ 1 u ̂ k ( x ) 0 and | u ̂ k 1 ( x ) u k ( x ) | λ k 1 . Then 2 k + 1 Ξ 1 u ̂ k 1 ( x ) 0 . We can find a real number α ( 0 , π 2 ) such that cos ( 2 α ) = ν ̂ k 1 , ν k at the point 𝑥, where the inner product is computed with respect to the metric 𝑔. Let us define a real number φ [ α , α ] by

    tan ( φ ) tan ( α ) = ( 1 + η ( λ k ( u ̂ k 1 u k ) ) ) | u ̂ k 1 | ( 1 η ( λ k ( u ̂ k 1 u k ) ) ) | u k | ( 1 + η ( λ k ( u ̂ k 1 u k ) ) ) | u ̂ k 1 | + ( 1 η ( λ k ( u ̂ k 1 u k ) ) ) | u k | ,

    where | u ̂ k 1 | and | u k | are computed with respect to the metric 𝑔. Then

    ν ̂ k = sin ( α + φ ) ν ̂ k 1 + sin ( α φ ) ν k sin ( 2 α )

    at the point 𝑥.

Proof

Statement (i) and statement (ii) follow directly from the definition.

To prove statement (iii), we consider an integer 1 k q and a point x Ω satisfying 2 k Ξ 1 u ̂ k ( x ) 0 and | u ̂ k 1 ( x ) u k ( x ) | λ k 1 . Using Lemma 2.6, we obtain

max { u ̂ k 1 ( x ) , u k ( x ) } u ̂ k ( x ) 0 , min { u ̂ k 1 ( x ) , u k ( x ) } max { u ̂ k 1 ( x ) , u k ( x ) } λ k 1 u ̂ k ( x ) 2 λ k 1 2 k + 1 Ξ 1 .

Therefore, 2 k + 1 Ξ 1 u ̂ k 1 ( x ) 0 and 2 k + 1 Ξ 1 u k ( x ) 0 . It follows from Lemma 2.11 that d u ̂ k 1 d u k 0 at the point 𝑥. Consequently, we can find a real number α ( 0 , π 2 ) such that cos ( 2 α ) = ν ̂ k 1 , ν k at the point 𝑥. We define a real number φ [ α , α ] so that

tan ( φ ) tan ( α ) = ( 1 + η ( λ k ( u ̂ k 1 u k ) ) ) | u ̂ k 1 | ( 1 η ( λ k ( u ̂ k 1 u k ) ) ) | u k | ( 1 + η ( λ k ( u ̂ k 1 u k ) ) ) | u ̂ k 1 | + ( 1 η ( λ k ( u ̂ k 1 u k ) ) ) | u k |

at the point 𝑥. Since d u ̂ k 1 d u k 0 at the point 𝑥, the denominator is strictly positive; this ensures that 𝜑 is well-defined.

We compute

sin ( α + φ ) ν ̂ k 1 + sin ( α φ ) ν k sin ( 2 α ) = 1 2 ( cos ( φ ) cos ( α ) + sin ( φ ) sin ( α ) ) ν ̂ k 1 + 1 2 ( cos ( φ ) cos ( α ) sin ( φ ) sin ( α ) ) ν k = 1 2 cos ( φ ) cos ( α ) ( 1 + tan ( φ ) tan ( α ) ) u ̂ k 1 | u ̂ k 1 | + 1 2 cos ( φ ) cos ( α ) ( 1 tan ( φ ) tan ( α ) ) u k | u k | = cos ( φ ) cos ( α ) ( 1 + η ( λ k ( u ̂ k 1 u k ) ) ) u ̂ k 1 + ( 1 η ( λ k ( u ̂ k 1 u k ) ) ) u k ( 1 + η ( λ k ( u ̂ k 1 u k ) ) ) | u ̂ k 1 | + ( 1 η ( λ k ( u ̂ k 1 u k ) ) ) | u k | = cos ( φ ) cos ( α ) 2 u ̂ k ( 1 + η ( λ k ( u ̂ k 1 u k ) ) ) | u ̂ k 1 | + ( 1 η ( λ k ( u ̂ k 1 u k ) ) ) | u k |

at the point 𝑥. The vector on the right-hand side is a nonnegative multiple of ν ̂ k . Using the identity

| sin ( α + φ ) ν ̂ k 1 + sin ( α φ ) ν k | 2 = sin 2 ( α + φ ) + sin 2 ( α φ ) + 2 sin ( α + φ ) sin ( α φ ) cos ( 2 α ) = sin 2 ( 2 α ) ,

we conclude that

sin ( α + φ ) ν ̂ k 1 + sin ( α φ ) ν k sin ( 2 α ) = ν ̂ k

at the point 𝑥. This completes the proof of Proposition 3.2. ∎

Proposition 3.3

If λ 0 is sufficiently large, we can find a collection of sets

W 0 , , W q R n

and a collection of maps N ̂ 0 : W 0 S n 1 , , N ̂ q : W q S n 1 with the following properties.

  1. For each 0 k q , W k is open and the map N ̂ k : W k S n 1 is smooth.

  2. We have W 0 = R n . Moreover, the map N ̂ 0 : W 0 S n 1 is constant and equal to N 0 .

  3. For each 1 k q , we have W k = ( W k 1 A k ) { u ̂ k 1 u k < λ k 1 } , where

    A k = W k 1 { | u ̂ k 1 u k | λ k 1 } ( { d u ̂ k 1 d u k = 0 } { N ̂ k 1 N k = 0 } ) .

  4. For each 0 k q , we have Ω { 2 k Ξ 1 u ̂ k 0 } W k .

  5. Let 1 k q . Suppose that 𝑥 is a point in W k satisfying u ̂ k 1 ( x ) u k ( x ) > λ k 1 . Then x W k 1 and N ̂ k = N ̂ k 1 at the point 𝑥.

  6. Let 1 k q . Suppose that 𝑥 is a point in W k satisfying u ̂ k 1 ( x ) u k ( x ) < λ k 1 . Then N ̂ k = N k at the point 𝑥.

  7. Let 1 k q . Suppose that 𝑥 is a point in W k satisfying | u ̂ k 1 ( x ) u k ( x ) | λ k 1 . Then x W k 1 A k . We can find real numbers α ( 0 , π 2 ) and θ ( 0 , π 2 α ) such that cos ( 2 α ) = ν ̂ k 1 , ν k and cos ( 2 θ α ) = N ̂ k 1 , N k at the point 𝑥. Let us define a real number φ [ α , α ] by

    tan ( φ ) tan ( α ) = ( 1 + η ( λ k ( u ̂ k 1 u k ) ) ) | u ̂ k 1 | ( 1 η ( λ k ( u ̂ k 1 u k ) ) ) | u k | ( 1 + η ( λ k ( u ̂ k 1 u k ) ) ) | u ̂ k 1 | + ( 1 η ( λ k ( u ̂ k 1 u k ) ) ) | u k | ,

    where | u ̂ k 1 | and | u k | are computed with respect to the metric 𝑔. Then

    N ̂ k = sin ( θ ( α + φ ) ) N ̂ k 1 + sin ( θ ( α φ ) ) N k sin ( 2 θ α )

    at the point 𝑥.

  8. Let 0 k q . Suppose that 𝑥 is a point in Ω { 2 k Ξ 1 u ̂ k 0 } . Then we can find nonnegative real numbers a 0 , , a k such that N ̂ k = i = 0 k a i N i at the point 𝑥. Moreover, a i = 0 for all 0 i k satisfying u i ( x ) < u ̂ k ( x ) 2 k λ i 1 .

Proof

We proceed by induction. We define W 0 = R n . Moreover, we define the map N ̂ 0 : W 0 S n 1 to be constant and equal to N 0 .

Suppose now that 1 k q , and that we have constructed sets W 0 , , W k 1 and maps N ̂ 0 : W 0 S n 1 , , N ̂ k 1 : W k 1 S n 1 satisfying properties (i)–(viii) above. We define

A k = W k 1 { | u ̂ k 1 u k | λ k 1 } ( { d u ̂ k 1 d u k = 0 } { N ̂ k 1 N k = 0 } ) , W k = ( W k 1 A k ) { u ̂ k 1 u k < λ k 1 } .

It follows from the induction hypothesis that W k 1 is open. Since A k is a relatively closed subset of W k 1 , it follows that W k is open.

It is clear that W k satisfies property (iii). Moreover, the induction hypothesis implies that

Ω { 2 k + 1 Ξ 1 u ̂ k 1 0 } W k 1 .

Using Lemma 2.6, we deduce that

Ω { 2 k Ξ 1 u ̂ k 0 } W k 1 { u ̂ k 1 u k < λ k 1 } .

On the other hand, using Lemma 2.6 and Lemma 2.11, we obtain d u ̂ k 1 d u k 0 at each point in Ω { 2 k Ξ 1 u ̂ k 0 } { | u ̂ k 1 u k | λ k 1 } . Moreover, since the map N ̂ k 1 satisfies property (viii), it follows from Lemma 2.4 and Lemma 2.6 that N ̂ k 1 N k 0 at each point in Ω { 2 k Ξ 1 u ̂ k 0 } { | u ̂ k 1 u k | λ k 1 } . Therefore, the set

Ω { 2 k Ξ 1 u ̂ k 0 }

is disjoint from A k . Putting these facts together, we conclude that

Ω { 2 k Ξ 1 u ̂ k 0 } W k .

Therefore, W k satisfies property (iv).

In the next step, we define the map N ̂ k : W k S n 1 . To that end, we distinguish three cases.

Case 1.  Suppose that 𝑥 is a point in W k satisfying u ̂ k 1 ( x ) u k ( x ) > λ k 1 . This implies x W k 1 . In this case, we define N ̂ k = N ̂ k 1 at the point 𝑥.

Case 2.  Suppose that 𝑥 is a point in W k satisfying u ̂ k 1 ( x ) u k ( x ) < λ k 1 . In this case, we define N ̂ k = N k at the point 𝑥.

Case 3.  Suppose that 𝑥 is a point in W k satisfying | u ̂ k 1 ( x ) u k ( x ) | λ k 1 . Then x W k 1 A k . This implies d u ̂ k 1 d u k 0 , and N ̂ k 1 N k 0 at the point 𝑥. Consequently, we can find real numbers α ( 0 , π 2 ) and θ ( 0 , π 2 α ) such that cos ( 2 α ) = ν ̂ k 1 , ν k and cos ( 2 θ α ) = N ̂ k 1 , N k at the point 𝑥. We define a real number φ [ α , α ] so that

tan ( φ ) tan ( α ) = ( 1 + η ( λ k ( u ̂ k 1 u k ) ) ) | u ̂ k 1 | ( 1 η ( λ k ( u ̂ k 1 u k ) ) ) | u k | ( 1 + η ( λ k ( u ̂ k 1 u k ) ) ) | u ̂ k 1 | + ( 1 η ( λ k ( u ̂ k 1 u k ) ) ) | u k |

at the point 𝑥. Since d u ̂ k 1 d u k 0 at the point 𝑥, the denominator is strictly positive; this ensures that 𝜑 is well-defined. With this understood, we define

N ̂ k = sin ( θ ( α + φ ) ) N ̂ k 1 + sin ( θ ( α φ ) ) N k sin ( 2 θ α )

at the point 𝑥. In view of the identity

| sin ( θ ( α + φ ) ) N ̂ k 1 + sin ( θ ( α φ ) ) N k | 2 = sin 2 ( θ ( α + φ ) ) + sin 2 ( θ ( α φ ) ) + 2 sin ( θ ( α + φ ) ) sin ( θ ( α φ ) ) cos ( 2 θ α ) = sin 2 ( 2 θ α ) ,

the map N ̂ k takes values in S n 1 .

It is clear from the definition that the map N ̂ k : W k S n 1 satisfies properties (v)–(vii). It remains to show that the map N ̂ k satisfies property (viii). To prove this, we again distinguish three cases.

Case 1.  Suppose that 𝑥 is a point in Ω with the property that 2 k Ξ 1 u ̂ k ( x ) 0 and u ̂ k 1 ( x ) u k ( x ) > λ k 1 . In this case, u ̂ k ( x ) = u ̂ k 1 ( x ) . In view of the induction hypothesis, we can find nonnegative real numbers b 0 , , b k 1 such that N ̂ k 1 = i = 0 k 1 b i N i at the point 𝑥. Moreover, b i = 0 for all 0 i k 1 satisfying u i ( x ) < u ̂ k 1 ( x ) 2 ( k 1 ) λ i 1 . We define nonnegative real numbers a 0 , , a k by a i = b i for 0 i k 1 and a k = 0 . Then

N ̂ k = N ̂ k 1 = i = 0 k 1 b i N i = i = 0 k a i N i

at the point 𝑥. Moreover, a i = 0 for all 0 i k satisfying u i ( x ) < u ̂ k ( x ) 2 k λ i 1 .

Case 2.  Suppose that 𝑥 is a point in Ω with the property that 2 k Ξ 1 u ̂ k ( x ) 0 and u ̂ k 1 ( x ) u k ( x ) < λ k 1 . In this case, u ̂ k ( x ) = u k ( x ) . We define nonnegative real numbers a 0 , , a k by a i = 0 for 0 i k 1 and a k = 1 . Then

N ̂ k = N k = i = 0 k a i N i .

Moreover, a i = 0 for all 0 i k satisfying u i ( x ) < u ̂ k ( x ) 2 k λ i 1 .

Case 3.  Suppose that 𝑥 is a point in Ω with the property that 2 k Ξ 1 u ̂ k ( x ) 0 and | u ̂ k 1 ( x ) u k ( x ) | λ k 1 . Using Lemma 2.6, we obtain

u ̂ k ( x ) max { u ̂ k 1 ( x ) , u k ( x ) } + λ k 1 min { u ̂ k 1 ( x ) , u k ( x ) } + 2 λ k 1 .

Let α ( 0 , π 2 ) , θ ( 0 , π 2 α ) , and φ [ α , α ] be defined as above. In view of the induction hypothesis, we can find nonnegative real numbers b 0 , , b k 1 such that N ̂ k 1 = i = 0 k 1 b i N i at the point 𝑥. Moreover, b i = 0 for all 0 i k 1 satisfying

u i ( x ) < u ̂ k 1 ( x ) 2 ( k 1 ) λ i 1 .

We define

a i = sin ( θ ( α + φ ) ) sin ( 2 θ α ) b i for 0 i k 1 and a k = sin ( θ ( α φ ) ) sin ( 2 θ α ) .

Since θ α ( 0 , π 2 ) and φ [ α , α ] , it follows that a 0 , , a k are nonnegative real numbers. Moreover,

N ̂ k = sin ( θ ( α + φ ) ) N ̂ k 1 + sin ( θ ( α φ ) ) N k sin ( 2 θ α ) = sin ( θ ( α + φ ) ) i = 0 k 1 b i N i + sin ( θ ( α φ ) ) N k sin ( 2 θ α ) = i = 0 k a i N i

at the point 𝑥. Finally, a i = 0 for all 0 i k satisfying u i ( x ) < u ̂ k ( x ) 2 k λ i 1 .

To summarize, we have shown that the map N ̂ k satisfies properties (v)–(viii). This completes the proof of Proposition 3.3. ∎

Remark 3.4

Let us explain the geometric significance of Proposition 3.2 (iii) and Proposition 3.3 (vii). Let 1 k q . Suppose that 𝑥 is a point in Ω with the property that

2 k Ξ 1 u ̂ k ( x ) 0 and | u ̂ k 1 ( x ) u k ( x ) | λ k 1 .

Let α ( 0 , π 2 ) , θ ( 0 , π 2 α ) , and φ [ α , α ] be defined as above. Then

( ν ̂ k 1 , ν k ) = 2 α , ( ν ̂ k 1 , ν ̂ k ) = α φ , and ( ν k , ν ̂ k ) = α + φ .

Moreover, ( N ̂ k 1 , N k ) = 2 θ α , ( N ̂ k 1 , N ̂ k ) = θ ( α φ ) , and ( N k , N ̂ k ) = θ ( α + φ ) .

Proposition 3.5

If λ 0 is sufficiently large, then the following holds. Let 0 j < k q . If 𝑥 is a point in Ω satisfying 2 j Ξ 1 u ̂ j ( x ) 0 and Ξ 1 u k ( x ) 0 , then

| N ̂ j N k | Ξ 1 .

Proof

Consider a point x Ω with the property that

2 j Ξ 1 u ̂ j ( x ) 0 and Ξ 1 u k ( x ) 0 .

By Proposition 3.3 (viii), we can find nonnegative real numbers a 0 , , a j such that

N ̂ j = i = 0 j a i N i

at the point 𝑥. Moreover, a i = 0 for all i { 0 , 1 , , j } satisfying u i ( x ) < u ̂ j ( x ) 2 j λ i 1 . Since

2 j Ξ 1 u ̂ j ( x ) 0 ,

it follows that a i = 0 for all i { 0 , 1 , , j } satisfying u i ( x ) 2 Ξ 1 . Using Lemma 2.4, we obtain

1 = | N ̂ j | i = 0 j a i Ξ | ( i = 0 j a i N i ) N k | = Ξ | N ̂ j N k |

at the point 𝑥. This completes the proof of Proposition 3.5. ∎

Proposition 3.6

Let 0 k q . Then | d N ̂ k | C λ k at each point in

Ω { 2 k Ξ 1 u ̂ k 0 } ,

where 𝐶 is independent of 𝛾 and λ 0 .

Proof

We argue by induction on 𝑘. Since the map N ̂ 0 is constant, the assertion is obvious for k = 0 .

Suppose now that 1 k q , and that | d N ̂ k 1 | C λ k 1 at each point in

Ω { 2 k + 1 Ξ 1 u ̂ k 1 0 } ,

where 𝐶 is independent of 𝛾 and λ 0 .

Let us consider an arbitrary point x 0 Ω { 2 k Ξ 1 u ̂ k 0 } . We distinguish three cases.

Case 1.  Suppose that u ̂ k 1 ( x 0 ) u k ( x 0 ) > λ k 1 . In this case, N ̂ k = N ̂ k 1 in an open neighborhood of the point x 0 . Therefore, | d N ̂ k | = | d N ̂ k 1 | C λ k at x 0 .

Case 2.  Suppose that u ̂ k 1 ( x 0 ) u k ( x 0 ) < λ k 1 . In this case, N ̂ k = N k in an open neighborhood of the point x 0 . Therefore, d N ̂ k = 0 at x 0 .

Case 3.  Suppose that | u ̂ k 1 ( x 0 ) u k ( x 0 ) | λ k 1 . In this case, 𝛼, 𝜃, and 𝜑 can be viewed as smooth functions which are defined in an open neighborhood of the point x 0 . Differentiating the identity cos ( 2 α ) = ν ̂ k 1 , ν k , we obtain | d cos ( 2 α ) | C λ k 1 at x 0 . By Lemma 2.11, α ( 0 , π 2 ) is bounded away from 0 and π 2 at the point x 0 . Consequently, we have | d α | C λ k 1 at x 0 . Differentiating the identity cos ( 2 θ α ) = N ̂ k 1 , N k , we obtain | d cos ( 2 θ α ) | C λ k 1 at x 0 . By Proposition 3.5, θ α ( 0 , π 2 ) is bounded away from 0 and π 2 at the point x 0 . Consequently, | d ( θ α ) | C λ k 1 at x 0 . This implies | d θ | C λ k 1 at x 0 . A similar argument gives | d φ | C λ k at x 0 . Since

N ̂ k = sin ( θ ( α + φ ) ) N ̂ k 1 + sin ( θ ( α φ ) ) N k sin ( 2 θ α )

in an open neighborhood of x 0 , we conclude that | d N ̂ k | C λ k at x 0 . This completes the proof of Proposition 3.6. ∎

Proposition 3.7

Suppose that 𝑔 is the Euclidean metric. For each 0 k q and each point x Ω { 2 k Ξ 1 u ̂ k 0 } , the vector N ̂ k ( x ) is the unit normal vector to the level set of u ̂ k passing through 𝑥.

Proof

This can be proved by induction on 𝑘, using the recursive relations in Proposition 3.2 and Proposition 3.3. ∎

Definition 3.8

We denote by N ̂ the restriction of the map N ̂ q to the hypersurface Σ ̂ = { u ̂ q = 0 } .

Proposition 3.9

The map N ̂ : Σ ̂ S n 1 is homotopic to the Gauss map of Σ ̂ .

Proof

In the special case when 𝑔 is the Euclidean metric, the assertion follows from Proposition 3.7. To prove the assertion in general, we deform the metric 𝑔 to the Euclidean metric. ∎

4 Controlling the angles under the smoothing procedure

Suppose that n 3 is an integer, and Ω = m = 0 q { u m 0 } is a compact, convex polytope in R n with non-empty interior. Throughout this section, we assume that Assumptions 1.1, 1.2, and 2.1 are satisfied. Moreover, we assume that 𝑔 is a Riemannian metric on R n which satisfies the following assumption.

Assumption 4.1

Let 0 j < k q . If 𝑥 is a point in Ω with u j ( x ) = u k ( x ) = 0 , then ν j , ν k N j , N k at the point 𝑥.

Lemma 4.2

Given ε ( 0 , 1 ) , we can find a small positive real number 𝛿 (depending on 𝜀) with the following property. Let 0 j < k q . If 𝑥 is a point in Ω with δ u j ( x ) 0 and δ u k ( x ) 0 , then ν j , ν k ε N j , N k at the point 𝑥.

Proof

Suppose that the assertion is false. We consider a sequence of counterexamples and pass to the limit. In the limit, we obtain a pair of integers 0 j < k q and a point x Ω such that u j ( x ) = u k ( x ) = 0 and ν j , ν k ε N j , N k at the point 𝑥. This contradicts Assumption 4.1. ∎

Lemma 4.3

We can find a constant Q > 1 (independent of 𝛾 and λ 0 ) such that the following holds. Consider a pair of integers 1 j < k q . Moreover, suppose that 𝑥 is a point in Ω satisfying 2 j Ξ 1 u ̂ j ( x ) 0 and | u ̂ j 1 ( x ) u j ( x ) | λ j 1 . If

ν ̂ j 1 , ν j ε N ̂ j 1 , N j , ν ̂ j 1 , ν k ε N ̂ j 1 , N k ε , ν j , ν k ε N j , N k ε

at the point 𝑥, then ν ̂ j , ν k Q ε N ̂ j , N k Q ε at the point 𝑥.

Proof

We can find real numbers α ( 0 , π 2 ) and θ ( 0 , π 2 α ) such that

cos ( 2 α ) = ν ̂ j 1 , ν j and cos ( 2 θ α ) = N ̂ j 1 , N j

at the point 𝑥. As above, we define a real number φ [ α , α ] by

tan ( φ ) tan ( α ) = ( 1 + η ( λ j ( u ̂ j 1 u j ) ) ) | u ̂ j 1 | ( 1 η ( λ j ( u ̂ j 1 u j ) ) ) | u j | ( 1 + η ( λ j ( u ̂ j 1 u j ) ) ) | u ̂ j 1 | + ( 1 η ( λ j ( u ̂ j 1 u j ) ) ) | u j | .

Using Proposition 3.2 (iii), we obtain

ν ̂ j = sin ( α + φ ) ν ̂ j 1 + sin ( α φ ) ν j sin ( 2 α )

at the point 𝑥. Moreover, Proposition 3.3 (vii) implies that

N ̂ j = sin ( θ ( α + φ ) ) N ̂ j 1 + sin ( θ ( α φ ) ) N j sin ( 2 θ α )

at the point 𝑥. This gives

(4.1) cos ( θ φ ) cos ( θ α ) ε N ̂ j , N k = sin ( θ ( α + φ ) ) sin ( 2 θ α ) ( ε N ̂ j 1 , N k ) + sin ( θ ( α φ ) ) sin ( 2 θ α ) ( ε N j , N k )

and

(4.2) N ̂ j , N k ν ̂ j , ν k + 2 cos ( φ ) cos ( α ) ε cos ( θ φ ) cos ( θ α ) ε = ( sin ( α + φ ) sin ( 2 α ) sin ( θ ( α + φ ) ) sin ( 2 θ α ) ) ( ε N ̂ j 1 , N k ) + ( sin ( α φ ) sin ( 2 α ) sin ( θ ( α φ ) ) sin ( 2 θ α ) ) ( ε N j , N k ) + sin ( α + φ ) sin ( 2 α ) ( N ̂ j 1 , N k ν ̂ j 1 , ν k + ε ) + sin ( α φ ) sin ( 2 α ) ( N j , N k ν j , ν k + ε )

at the point 𝑥. By assumption,

ε N ̂ j 1 , N k 0 , ε N j , N k 0 , N ̂ j 1 , N k ν ̂ j 1 , ν k + ε 0 , N j , N k ν j , ν k + ε 0

at the point 𝑥. Using identity (4.1), we obtain

(4.3) cos ( θ φ ) cos ( θ α ) ε N ̂ j , N k 0

at the point 𝑥.

In the next step, we bound N ̂ j , N k ν ̂ j , ν k from below. To that end, we distinguish two cases.

Case 1.  Suppose that θ ( 0 , 1 ) . Applying Proposition B.2 with β = α + φ 2 [ 0 , α ] gives

sin ( α + φ ) sin ( 2 α ) sin ( θ ( α + φ ) ) sin ( 2 θ α ) 0 .

Similarly, applying Proposition B.2 with β = α φ 2 [ 0 , α ] gives

sin ( α φ ) sin ( 2 α ) sin ( θ ( α φ ) ) sin ( 2 θ α ) 0 .

Using identity (4.2), we conclude that

(4.4) N ̂ j , N k ν ̂ j , ν k + 2 cos ( φ ) cos ( α ) ε cos ( θ φ ) cos ( θ α ) ε 0

at the point 𝑥.

Case 2.  Suppose that θ [ 1 , π 2 α ) . Applying Proposition B.3 with β = α + φ 2 [ 0 , α ] gives

sin ( α + φ ) sin ( 2 α ) sin ( θ ( α + φ ) ) sin ( 2 θ α ) 4 ( θ 1 ) α sin ( 2 α ) sin ( 2 θ α ) .

Similarly, applying Proposition B.3 with β = α φ 2 [ 0 , α ] gives

sin ( α φ ) sin ( 2 α ) sin ( θ ( α φ ) ) sin ( 2 θ α ) 4 ( θ 1 ) α sin ( 2 α ) sin ( 2 θ α ) .

Using identity (4.2), we obtain

N ̂ j , N k ν ̂ j , ν k + 2 cos ( φ ) cos ( α ) ε cos ( θ φ ) cos ( θ α ) ε 4 ( θ 1 ) α sin ( 2 α ) sin ( 2 θ α ) ( 2 ε N ̂ j 1 , N k N j , N k ) 8 ( θ 1 ) α ( 1 + ε ) sin ( 2 α ) sin ( 2 θ α )

at the point 𝑥. On the other hand,

2 ( θ 1 ) α inf t [ 2 α , 2 θ α ] sin ( t ) cos ( 2 α ) cos ( 2 θ α ) = ν ̂ j 1 , ν j N ̂ j 1 , N j ε

at the point 𝑥. Putting these facts together, we conclude that

(4.5) N ̂ j , N k ν ̂ j , ν k + 2 cos ( φ ) cos ( α ) ε cos ( θ φ ) cos ( θ α ) ε 4 ε ( 1 + ε ) sin ( 2 α ) sin ( 2 θ α ) inf t [ 2 α , 2 θ α ] sin ( t )

at the point 𝑥.

By Lemma 2.11, α ( 0 , π 2 ) is bounded away from 0 and π 2 at the point 𝑥. By Proposition 3.5, θ α ( 0 , π 2 ) is bounded away from 0 and π 2 at the point 𝑥. Combining (4.3), (4.4), and (4.5), the assertion follows. This completes the proof of Lemma 4.3. ∎

Proposition 4.4

Let 0 j q 1 , and let ε ( 0 , 1 ) be given. Then we can find a real number δ ( 0 , 2 j Ξ 1 ) (depending on 𝜀, but independent of 𝛾 and λ 0 ) with the following property. Suppose that λ 0 2 δ 1 , k { j + 1 , , q } , and 𝑥 is a point in Ω satisfying δ u ̂ j ( x ) 0 and δ u k ( x ) 0 . Then ν ̂ j , ν k ε N ̂ j , N k ε at the point 𝑥.

Proof

We argue by induction on 𝑗. For j = 0 , the assertion follows from Lemma 2.2 and Lemma 4.2. We next assume that 1 j q , and that the assertion is true for j 1 . We claim that the assertion is true for 𝑗. To prove this, let us fix a constant Q > 1 so that the conclusion of Lemma 4.3 holds. Note that 𝑄 is independent of 𝛾 and λ 0 . Let ε ( 0 , 1 ) be given. In view of the induction hypothesis, we can find a real number δ 0 ( 0 , 2 j + 1 Ξ 1 ) (depending on 𝜀, but independent of 𝛾 and λ 0 ) such that

(4.6) ν ̂ j 1 , ν k Q 1 ε N ̂ j 1 , N k Q 1 ε

whenever λ 0 2 δ 0 1 , k { j , , q } , and 𝑥 is a point in Ω satisfying δ 0 u ̂ j 1 ( x ) 0 and δ 0 u k ( x ) 0 . By Lemma 2.2 and Lemma 4.2, we can find a positive real number δ 1 (depending on 𝜀) such that

(4.7) ν j , ν k Q 1 ε N j , N k Q 1 ε

whenever k { j + 1 , , q } and 𝑥 is a point in Ω satisfying

δ 1 u j ( x ) 0 and δ 1 u k ( x ) 0 .

We define δ : = 1 2 min { δ 0 , δ 1 } ( 0 , 2 j Ξ 1 ) . Note that 𝛿 may depend on 𝜀, but is independent of 𝛾 and λ 0 . We claim that 𝛿 has the desired property. To prove this, we assume that λ 0 2 δ 1 . Moreover, we consider an integer k { j + 1 , , q } and a point x Ω such that δ u ̂ j ( x ) 0 and δ u k ( x ) 0 . We distinguish three cases.

Case 1.  Suppose first that u ̂ j 1 ( x ) u j ( x ) > λ j 1 . In this case, u ̂ j ( x ) = u ̂ j 1 ( x ) . Consequently, δ u ̂ j 1 ( x ) 0 and δ u k ( x ) 0 . Using (4.6), we obtain

ν ̂ j 1 , ν k Q 1 ε N ̂ j 1 , N k Q 1 ε

at the point 𝑥. Since ν ̂ j = ν ̂ j 1 and N ̂ j = N ̂ j 1 at the point 𝑥, it follows that

ν ̂ j , ν k Q 1 ε N ̂ j , N k Q 1 ε

at the point 𝑥.

Case 2.  Suppose next that u ̂ j 1 ( x ) u j ( x ) < λ j 1 . In this case, u ̂ j ( x ) = u j ( x ) . Consequently, δ u j ( x ) 0 and δ u k ( x ) 0 . Using (4.7), we obtain

ν j , ν k Q 1 ε N j , N k Q 1 ε .

Since ν ̂ j = ν j and N ̂ j = N j at the point 𝑥, it follows that

ν ̂ j , ν k Q 1 ε N ̂ j , N k Q 1 ε

at the point 𝑥.

Case 3.  Suppose finally that | u ̂ j 1 ( x ) u j ( x ) | λ j 1 . As u ̂ j ( x ) [ δ , 0 ] , Lemma 2.6 implies that u ̂ j 1 ( x ) [ δ 2 λ j 1 , 0 ] and u j ( x ) [ δ 2 λ j 1 , 0 ] . Since λ j λ 0 2 δ 1 , we conclude that 2 δ u ̂ j 1 ( x ) 0 and 2 δ u j ( x ) 0 . Moreover, δ u k ( x ) 0 . Using (4.6), we obtain

ν ̂ j 1 , ν j Q 1 ε N ̂ j 1 , N j Q 1 ε , ν ̂ j 1 , ν k Q 1 ε N ̂ j 1 , N k Q 1 ε

at the point 𝑥. Using (4.7), we obtain

ν j , ν k Q 1 ε N j , N k Q 1 ε

at the point 𝑥. Using Lemma 4.3, we conclude that

ν ̂ j , ν k ε N ̂ j , N k ε

at the point 𝑥. This completes the proof of Proposition 4.4. ∎

5 Estimating max { d N ̂ tr H , 0 }

Suppose that n 3 is an integer, and Ω = m = 0 q { u m 0 } is a compact, convex polytope in R n with non-empty interior. Throughout this section, we assume that Assumptions 1.1, 1.2, and 2.1 are satisfied. Moreover, we assume that 𝑔 is a Riemannian metric on R n which satisfies Assumption 4.1. We further assume that the metric 𝑔 satisfies the following assumption.

Assumption 5.1

For each k { 0 , 1 , , q } , the mean curvature of the hypersurface { u k = 0 } with respect to 𝑔 is nonnegative at each point in Ω { u k = 0 } .

In the remainder of this section, we will estimate the quantity max { d N ̂ tr H , 0 } , where 𝐻 denotes the mean curvature of Σ ̂ with respect to the metric 𝑔.

Proposition 5.2

Let 0 k q . Then H d N ̂ tr 0 on the set F k .

Proof

Let us fix a point x 0 F k . We consider a small open neighborhood 𝑊 of x 0 . We choose 𝑊 so that

W m = k + 1 q { u ̂ m 1 u m > λ m 1 } .

If k 0 , we further require that W { u ̂ k 1 u k < λ k 1 } .

Then u k = u ̂ k = u ̂ q at each point in 𝑊. This implies ν k = ν ̂ k = ν ̂ q = ν ̂ at each point in Σ ̂ W . Moreover, it follows from Proposition 3.3 (v) and (vi) that N k = N ̂ k = N ̂ q = N ̂ at each point in Σ ̂ W . Using Assumption 5.1, we conclude that H 0 and d N ̂ tr = 0 at x 0 . This completes the proof of Proposition 5.2. ∎

Proposition 5.3

Let 0 j < k q . If λ 0 is sufficiently large, then

H d N ̂ tr C λ k max { ν j , ν k N j , N k , 0 } C

on the set E j , k . Here, 𝐶 is independent of 𝛾 and λ 0 .

Proof

Let us fix a point x 0 E j , k . In the following, we will consider a small open neighborhood 𝑊 of x 0 . We choose 𝑊 so that

W m = k + 1 q { u ̂ m 1 u m > λ m 1 } m = j + 1 k 1 { u ̂ m 1 u m > λ m 1 } .

If j 0 , we further require that W { u ̂ j 1 u j < λ j 1 } .

Then u j = u ̂ j = u ̂ k 1 and u ̂ k = u ̂ q at each point in 𝑊. This implies ν j = ν ̂ j = ν ̂ k 1 and ν ̂ k = ν ̂ q = ν ̂ at each point in Σ ̂ W . Moreover, it follows from Proposition 3.3 (v) and (vi) that N j = N ̂ j = N ̂ k 1 and N ̂ k = N ̂ q = N ̂ at each point in Σ ̂ W .

We define a smooth function α : Σ ̂ W ( 0 , π 2 ) by

(5.1) cos ( 2 α ) = ν j , ν k

at each point in Σ ̂ W , and we define a smooth function θ : Σ ̂ W R so that θ ( 0 , π 2 α ) at each point in Σ ̂ W and

(5.2) cos ( 2 θ α ) = N j , N k

at each point in Σ ̂ W . Finally, we define a function φ : Σ ̂ W R so that φ [ α , α ] at each point in Σ ̂ W and

(5.3) tan ( φ ) tan ( α ) = ( 1 + η ( λ k ( u j u k ) ) ) | u j | ( 1 η ( λ k ( u j u k ) ) ) | u k | ( 1 + η ( λ k ( u j u k ) ) ) | u j | + ( 1 η ( λ k ( u j u k ) ) ) | u k |

at each point in Σ ̂ W , where | u j | and | u k | are computed with respect to the metric 𝑔.

Differentiating identity (5.1) gives | Σ ̂ cos ( 2 α ) | C at x 0 , where 𝐶 is independent of 𝛾 and λ 0 . Since α ( 0 , π 2 ) is bounded away from 0 and π 2 at the point x 0 , it follows that | Σ ̂ α | C at x 0 , where 𝐶 is independent of 𝛾 and λ 0 . Differentiating identity (5.2) gives Σ ̂ cos ( 2 θ α ) = 0 at x 0 . Since θ α ( 0 , π 2 ) , it follows that Σ ̂ ( θ α ) = 0 at x 0 . This implies | Σ ̂ θ | C at x 0 , where 𝐶 is independent of 𝛾 and λ 0 . Finally, differentiating identity (5.3) gives

| Σ ̂ ( tan ( φ ) tan ( α ) ) λ k Υ ( Σ ̂ u j Σ ̂ u k ) | C

at x 0 . Here, 𝐶 is independent of 𝛾 and λ 0 , and Υ is defined by

Υ = 4 η ′′ ( λ k ( u j u k ) ) | u j | | u k | ( ( 1 + η ( λ k ( u j u k ) ) ) | u j | + ( 1 η ( λ k ( u j u k ) ) ) | u k | ) 2 .

Note that 0 Υ C at the point x 0 , where 𝐶 is independent of 𝛾 and λ 0 . This implies

(5.4) | Σ ̂ φ λ k Υ tan ( α ) cos 2 ( φ ) ( Σ ̂ u j Σ ̂ u k ) | C

at the point x 0 , where 𝐶 is independent of 𝛾 and λ 0 .

By Proposition 3.2 (iii), the unit normal to Σ ̂ is given by

(5.5) ν ̂ = sin ( α + φ ) ν j + sin ( α φ ) ν k sin ( 2 α )

at each point in Σ ̂ W . Differentiating identity (5.5) gives

(5.6) | D ξ ν ̂ d φ ( ξ ) cos ( α + φ ) ν j cos ( α φ ) ν k sin ( 2 α ) | C | ξ |

at x 0 , where 𝜉 denotes an arbitrary vector in T x 0 Σ ̂ and 𝐶 is independent of 𝛾 and λ 0 . In view of the identities

cos ( α + φ ) ν j cos ( α φ ) ν k , sin ( α + φ ) ν j + sin ( α φ ) ν k = cos ( α + φ ) sin ( α + φ ) cos ( α φ ) sin ( α φ ) + cos ( α + φ ) sin ( α φ ) cos ( 2 α ) cos ( α φ ) sin ( α + φ ) cos ( 2 α ) = 0

and

| cos ( α + φ ) ν j cos ( α φ ) ν k | 2 = cos 2 ( α + φ ) + cos 2 ( α φ ) 2 cos ( α + φ ) cos ( α φ ) cos ( 2 α ) = sin 2 ( 2 α ) ,

the vector

cos ( α + φ ) ν j cos ( α φ ) ν k sin ( 2 α )

lies in the tangent space T x 0 Σ ̂ and has unit length with respect to the metric 𝑔. Estimate (5.6) implies

(5.7) H Σ ̂ φ , cos ( α + φ ) ν j cos ( α φ ) ν k sin ( 2 α ) C

at x 0 , where 𝐶 is independent of 𝛾 and λ 0 . Combining (5.4) and (5.7), we obtain

(5.8) H λ k Υ tan ( α ) cos 2 ( φ ) Σ ̂ u j Σ ̂ u k , cos ( α + φ ) ν j cos ( α φ ) ν k sin ( 2 α ) C

at x 0 , where 𝐶 is independent of 𝛾 and λ 0 . Using identity (5.5), we obtain

ν j = sin ( α φ ) cos ( α + φ ) ν j cos ( α φ ) ν k sin ( 2 α ) + cos ( α φ ) ν ̂ , ν k = sin ( α + φ ) cos ( α + φ ) ν j cos ( α φ ) ν k sin ( 2 α ) + cos ( α + φ ) ν ̂ .

This implies

u j u k = | u j | ν j | u k | ν k = ( sin ( α φ ) | u j | + sin ( α + φ ) | u k | ) cos ( α + φ ) ν j cos ( α φ ) ν k sin ( 2 α ) + ( cos ( α φ ) | u j | cos ( α + φ ) | u k | ) ν ̂ .

Projecting to T x 0 Σ ̂ gives

Σ ̂ u j Σ ̂ u k = ( sin ( α φ ) | u j | + sin ( α + φ ) | u k | ) cos ( α + φ ) ν j cos ( α φ ) ν k sin ( 2 α ) .

Since sin ( α φ ) | u j | + sin ( α + φ ) | u k | 0 , it follows that

Σ ̂ u j Σ ̂ u k , cos ( α + φ ) ν j cos ( α φ ) ν k sin ( 2 α ) = | Σ ̂ u j Σ ̂ u k | .

Substituting this identity into (5.8) gives

(5.9) H λ k Υ tan ( α ) cos 2 ( φ ) | Σ ̂ u j Σ ̂ u k | C

at x 0 , where 𝐶 is independent of 𝛾 and λ 0 .

On the other hand, it follows from Proposition 3.3 (vii) that

(5.10) N ̂ = sin ( θ ( α + φ ) ) N j + sin ( θ ( α φ ) ) N k sin ( 2 θ α )

at each point in Σ ̂ W . Differentiating identity (5.10) gives

(5.11) | d N ̂ ( ξ ) θ d φ ( ξ ) cos ( θ ( α + φ ) ) N j cos ( θ ( α φ ) ) N k sin ( 2 θ α ) | C | ξ |

at x 0 , where 𝜉 denotes an arbitrary vector in T x 0 Σ ̂ and 𝐶 is independent of 𝛾 and λ 0 . In view of the identity

| cos ( θ ( α + φ ) ) N j cos ( θ ( α φ ) ) N k | 2 = cos 2 ( θ ( α + φ ) ) + cos 2 ( θ ( α φ ) ) 2 cos ( θ ( α + φ ) ) cos ( θ ( α φ ) ) cos ( 2 θ α ) = sin 2 ( 2 θ α ) ,

the vector

cos ( θ ( α + φ ) ) N j cos ( θ ( α φ ) ) N k sin ( 2 θ α ) R n

has unit length with respect to the Euclidean metric. Therefore, estimate (5.11) implies

(5.12) d N ̂ tr θ | Σ ̂ φ | + C

at x 0 , where 𝐶 is independent of 𝛾 and λ 0 . Combining (5.4) and (5.12), we obtain

(5.13) d N ̂ tr θ λ k Υ tan ( α ) cos 2 ( φ ) | Σ ̂ u j Σ ̂ u k | + C

at x 0 , where 𝐶 is independent of 𝛾 and λ 0 . Combining (5.9) and (5.13), we conclude that

H d N ̂ tr ( 1 θ ) λ k Υ tan ( α ) cos 2 ( φ ) | Σ ̂ u j Σ ̂ u k | C C λ k max { θ 1 , 0 } C

at x 0 , where 𝐶 is independent of 𝛾 and λ 0 . This completes the proof of Proposition 5.3. ∎

Corollary 5.4

Let 0 j < k q . If λ 0 is sufficiently large (depending on 𝛾), then H d N ̂ tr C γ λ k on the set E j , k . Here, 𝐶 is independent of 𝛾 and λ 0 .

Proof

We apply Lemma 4.2 with ε = γ . Consequently, if λ 0 is sufficiently large (depending on 𝛾), then ν j , ν k N j , N k γ at each point in

Ω { 2 λ 0 1 u j 0 } { 2 λ 0 1 u k 0 } .

Since

E j , k Ω { 2 λ k 1 u j 0 } { 2 λ k 1 u k 0 } ,

it follows that ν j , ν k N j , N k γ at each point in E j , k . The assertion follows now from Proposition 5.3. This completes the proof of Corollary 5.4. ∎

Corollary 5.5

We can find a constant M > 2 (independent of 𝛾 and λ 0 ) with the following property. Let 0 j < k q . If λ 0 is sufficiently large, then H d N ̂ tr C on the set

E j , k i { 0 , 1 , , q } { j , k } { M λ k 1 u i 0 } .

Here, 𝐶 is independent of 𝛾 and λ 0 .

Proof

We distinguish two cases.

Case 1.  Suppose that Ω { u j = 0 } { u k = 0 } = . Then

Ω { 2 λ 0 1 u k 0 } { 2 λ 0 1 u j 0 } =

if λ 0 is sufficiently large. Consequently, E j , k = if λ 0 is sufficiently large. Thus, the assertion is trivially true in this case.

Case 2.  Suppose that Ω { u j = 0 } { u k = 0 } . It follows from Assumption 1.1 that the hyperplanes { u j = 0 } and { u k = 0 } must intersect transversally. Let us consider a point x E j , k . Then 2 λ k 1 u j ( x ) 0 and 2 λ k 1 u k ( x ) 0 . By transversality, we can find a point 𝑦 such that u j ( y ) = u k ( y ) = 0 and d eucl ( x , y ) M λ k 1 , where M > 2 is independent of 𝛾 and λ 0 .

Suppose now that 𝑥 is a point in E j , k i { 0 , 1 , , q } { j , k } { M λ k 1 u i 0 } . Then u i ( x ) < M λ k 1 for all i { 0 , 1 , , q } { j , k } . As | u i ( x ) u i ( y ) | d eucl ( x , y ) M λ k 1 for all i { 0 , 1 , , q } , it follows that u i ( y ) < 0 for all i { 0 , 1 , , q } { j , k } , and consequently y Ω . Assumption 4.1 implies that ν j , ν k N j , N k 0 at the point 𝑦. Since d eucl ( x , y ) M λ k 1 , it follows that ν j , ν k N j , N k C λ k 1 at the point 𝑥, where 𝐶 is independent of 𝛾 and λ 0 . Using Proposition 5.3, we conclude that H d N ̂ tr C at the point 𝑥, where 𝐶 is independent of 𝛾 and λ 0 . This completes the proof of Corollary 5.5. ∎

Proposition 5.6

Let 0 i < j < k q . If λ 0 is sufficiently large, then

H d N ̂ tr C λ k max { ν ̂ j , ν k N ̂ j , N k , 0 } C λ j

on the set G i , j , k . Here, 𝐶 is independent of 𝛾 and λ 0 .

Proof

Let us fix a point x 0 G i , j , k . In the following, we will consider a small open neighborhood 𝑊 of x 0 . We choose 𝑊 so that

W m = k + 1 q { u ̂ m 1 u m > λ m 1 } m = j + 1 k 1 { u ̂ m 1 u m > λ m 1 } .

Then u ̂ j = u ̂ k 1 and u ̂ k = u ̂ q at each point in 𝑊. This implies ν ̂ j = ν ̂ k 1 and ν ̂ k = ν ̂ q = ν ̂ at each point in Σ ̂ W . Moreover, it follows from Proposition 3.3 (v) that N ̂ j = N ̂ k 1 and N ̂ k = N ̂ q = N ̂ at each point in Σ ̂ W .

We define a smooth function α : Σ ̂ W ( 0 , π 2 ) by

(5.14) cos ( 2 α ) = ν ̂ j , ν k

at each point in Σ ̂ W , and we define a smooth function θ : Σ ̂ W R so that θ ( 0 , π 2 α ) at each point in Σ ̂ W and

(5.15) cos ( 2 θ α ) = N ̂ j , N k

at each point in Σ ̂ W . Finally, we define a function φ : Σ ̂ W R so that φ [ α , α ] at each point in Σ ̂ W and

(5.16) tan ( φ ) tan ( α ) = ( 1 + η ( λ k ( u ̂ j u k ) ) ) | u ̂ j | ( 1 η ( λ k ( u ̂ j u k ) ) ) | u k | ( 1 + η ( λ k ( u ̂ j u k ) ) ) | u ̂ j | + ( 1 η ( λ k ( u ̂ j u k ) ) ) | u k |

at each point in Σ ̂ W , where | u ̂ j | and | u k | are computed with respect to the metric 𝑔.

Let us differentiate identity (5.14). Using Lemma 2.8 and Lemma 2.10, we obtain

| Σ ̂ cos ( 2 α ) | C λ j

at x 0 , where 𝐶 is independent of 𝛾 and λ 0 . Since α ( 0 , π 2 ) is bounded away from 0 and π 2 at the point x 0 , it follows that | Σ ̂ α | C λ j at x 0 , where 𝐶 is independent of 𝛾 and λ 0 . We next differentiate identity (5.15). Using Proposition 3.6, we obtain | Σ ̂ cos ( 2 θ α ) | C λ j at x 0 . Since θ α ( 0 , π 2 ) is bounded away from 0 and π 2 at the point x 0 , it follows that | Σ ̂ ( θ α ) | C λ j at x 0 , where 𝐶 is independent of 𝛾 and λ 0 . This implies | Σ ̂ θ | C λ j at x 0 , where 𝐶 is independent of 𝛾 and λ 0 . Finally, we differentiate identity (5.16). Using Lemma 2.8 and Lemma 2.10, we obtain

| Σ ̂ ( tan ( φ ) tan ( α ) ) λ k Υ ( Σ ̂ u ̂ j Σ ̂ u k ) | C λ j

at x 0 . Here, 𝐶 is independent of 𝛾 and λ 0 , and Υ is defined by

Υ = 4 η ′′ ( λ k ( u ̂ j u k ) ) | u ̂ j | | u k | ( ( 1 + η ( λ k ( u ̂ j u k ) ) ) | u ̂ j | + ( 1 η ( λ k ( u ̂ j u k ) ) ) | u k | ) 2 .

Note that 0 Υ C at the point x 0 , where 𝐶 is independent of 𝛾 and λ 0 . This implies

(5.17) | Σ ̂ φ λ k Υ tan ( α ) cos 2 ( φ ) ( Σ ̂ u ̂ j Σ ̂ u k ) | C λ j

at the point x 0 , where 𝐶 is independent of 𝛾 and λ 0 .

By Proposition 3.2 (iii), the unit normal to Σ ̂ is given by

(5.18) ν ̂ = sin ( α + φ ) ν ̂ j + sin ( α φ ) ν k sin ( 2 α )

at each point in Σ ̂ W . Differentiating identity (5.18) gives

(5.19) | D ξ ν ̂ d φ ( ξ ) cos ( α + φ ) ν ̂ j cos ( α φ ) ν k sin ( 2 α ) | C λ j | ξ |

at x 0 , where 𝜉 denotes an arbitrary vector in T x 0 Σ ̂ and 𝐶 is independent of 𝛾 and λ 0 . The vector

cos ( α + φ ) ν ̂ j cos ( α φ ) ν k sin ( 2 α )

lies in the tangent space T x 0 Σ ̂ and has unit length with respect to the metric 𝑔. Estimate (5.19) implies

(5.20) H Σ ̂ φ , cos ( α + φ ) ν ̂ j cos ( α φ ) ν k sin ( 2 α ) C λ j

at x 0 , where 𝐶 is independent of 𝛾 and λ 0 . Combining (5.17) and (5.20), we obtain

(5.21) H λ k Υ tan ( α ) cos 2 ( φ ) Σ ̂ u ̂ j Σ ̂ u k , cos ( α + φ ) ν ̂ j cos ( α φ ) ν k sin ( 2 α ) C λ j

at x 0 , where 𝐶 is independent of 𝛾 and λ 0 . Using identity (5.18), we obtain

ν ̂ j = sin ( α φ ) cos ( α + φ ) ν ̂ j cos ( α φ ) ν k sin ( 2 α ) + cos ( α φ ) ν ̂ , ν k = sin ( α + φ ) cos ( α + φ ) ν ̂ j cos ( α φ ) ν k sin ( 2 α ) + cos ( α + φ ) ν ̂ .

This implies

u ̂ j u k = | u ̂ j | ν ̂ j | u k | ν k = ( sin ( α φ ) | u ̂ j | + sin ( α + φ ) | u k | ) cos ( α + φ ) ν ̂ j cos ( α φ ) ν k sin ( 2 α ) + ( cos ( α φ ) | u ̂ j | cos ( α + φ ) | u k | ) ν ̂ .

Projecting to T x 0 Σ ̂ gives

Σ ̂ u ̂ j Σ ̂ u k = ( sin ( α φ ) | u ̂ j | + sin ( α + φ ) | u k | ) cos ( α + φ ) ν ̂ j cos ( α φ ) ν k sin ( 2 α ) .

Since sin ( α φ ) | u ̂ j | + sin ( α + φ ) | u k | 0 , it follows that

Σ ̂ u ̂ j Σ ̂ u k , cos ( α + φ ) ν ̂ j cos ( α φ ) ν k sin ( 2 α ) = | Σ ̂ u ̂ j Σ ̂ u k | .

Substituting this identity into (5.21) gives

(5.22) H λ k Υ tan ( α ) cos 2 ( φ ) | Σ ̂ u ̂ j Σ ̂ u k | C λ j

at x 0 , where 𝐶 is independent of 𝛾 and λ 0 .

On the other hand, it follows from Proposition 3.3 (vii) that

(5.23) N ̂ = sin ( θ ( α + φ ) ) N ̂ j + sin ( θ ( α φ ) ) N k sin ( 2 θ α )

at each point in Σ ̂ W . Differentiating identity (5.23) gives

(5.24) | d N ̂ ( ξ ) θ d φ ( ξ ) cos ( θ ( α + φ ) ) N ̂ j cos ( θ ( α φ ) ) N k sin ( 2 θ α ) | C λ j | ξ |

at x 0 , where 𝜉 denotes an arbitrary vector in T x 0 Σ ̂ and 𝐶 is independent of 𝛾 and λ 0 . The vector

cos ( θ ( α + φ ) ) N ̂ j cos ( θ ( α φ ) ) N k sin ( 2 θ α ) R n

has unit length with respect to the Euclidean metric. Therefore, estimate (5.24) implies

(5.25) d N ̂ tr θ | Σ ̂ φ | + C λ j

at x 0 , where 𝐶 is independent of 𝛾 and λ 0 . Combining (5.17) and (5.25), we obtain

(5.26) d N ̂ tr θ λ k Υ tan ( α ) cos 2 ( φ ) | Σ ̂ u ̂ j Σ ̂ u k | + C λ j

at x 0 , where 𝐶 is independent of 𝛾 and λ 0 . Combining (5.22) and (5.26), we conclude that

H d N ̂ tr ( 1 θ ) λ k Υ tan ( α ) cos 2 ( φ ) | Σ ̂ u ̂ j Σ ̂ u k | C λ j C λ k max { θ 1 , 0 } C λ j

at x 0 , where 𝐶 is independent of 𝛾 and λ 0 . This completes the proof of Proposition 5.6. ∎

Corollary 5.7

Let 0 i < j < k q . If λ 0 is sufficiently large (depending on 𝛾), then H d N ̂ tr C γ λ k on the set G i , j , k . Here, 𝐶 is independent of 𝛾 and λ 0 .

Proof

We apply Proposition 4.4 with ε = γ . Consequently, if λ 0 is sufficiently large (depending on 𝛾), then ν ̂ j , ν k N ̂ j , N k γ at each point in

Ω { 2 λ 0 1 u ̂ j 0 } { 2 λ 0 1 u k 0 } .

Since

G i , j , k Ω { 2 λ k 1 u ̂ j 0 } { 2 λ k 1 u k 0 } ,

it follows that ν ̂ j , ν k N ̂ j , N k γ at each point in G i , j , k . The assertion follows now from Proposition 5.6. This completes the proof of Corollary 5.7. ∎

In the remainder of this section, we will estimate a suitable Morrey norm of the function max { d N ̂ tr H , 0 } on Σ ̂ . To that end, we will combine Proposition 5.2, Corollary 5.4, Corollary 5.5, and Corollary 5.7 with the area estimates in Lemma 2.12 and Lemma 2.13.

Proposition 5.8

Fix an exponent σ ( 1 , q q 1 ) . Suppose γ ( 0 , M 1 ) , where M > 2 denotes the constant in Corollary 5.5. If λ 0 is sufficiently large (depending on 𝛾), then

r σ + 1 n E j , k B r ( p ) ( max { d N ̂ tr H , 0 } ) σ C γ σ q ( σ 1 )

for all 0 j < k q and all 0 < r 1 . Here, 𝐶 is independent of 𝛾 and λ 0 .

Proof

If λ 0 is sufficiently large (depending on 𝛾), then Corollary 5.4 implies that max { d N ̂ tr H , 0 } C γ λ k on the set E j , k , where 𝐶 is independent of 𝛾 and λ 0 . Moreover, Corollary 5.5 implies that max { d N ̂ tr H , 0 } C on the set

E j , k i { 0 , 1 , , q } { j , k } { M λ k 1 u i 0 } ,

where, again, 𝐶 is independent of 𝛾 and λ 0 . We distinguish two cases.

Case 1.  Suppose that 0 < r λ 0 1 . Since

E j , k Ω { u ̂ k = 0 } { 2 λ k 1 u ̂ k 1 0 } { 2 λ k 1 u k 0 } ,

Lemma 2.12 implies that | E j , k B r ( p ) | C λ k 1 r n 2 , where 𝐶 is independent of 𝛾 and λ 0 . Consequently,

r σ + 1 n E j , k B r ( p ) ( max { d N ̂ tr H , 0 } ) σ C r σ + 1 n ( γ λ k ) σ | E j , k B r ( p ) | C γ σ ( λ k r ) σ 1 C γ σ q ( σ 1 ) ,

where 𝐶 is independent of 𝛾 and λ 0 .

Case 2.  Suppose that λ 0 1 r 1 . As above, | E j , k B r ( p ) | C λ k 1 r n 2 by Lemma 2.12. Moreover, since γ ( 0 , M 1 ) , we have M λ k 1 ( γ λ k ) 1 λ 0 1 . This implies

E j , k { M λ k 1 u i 0 } Ω { u ̂ k = 0 } { 2 λ k 1 u ̂ k 1 0 } { 2 λ k 1 u k 0 } { 4 λ 0 1 u j 0 } { 6 λ 0 1 u i 0 }

for each i { 0 , 1 , , q } { j , k } . Using Lemma 2.13, we obtain

| E j , k { M λ k 1 u i 0 } B r ( p ) | C λ k 1 λ 0 1 r n 3

for each i { 0 , 1 , , q } { j , k } , where 𝐶 is independent of 𝛾 and λ 0 . Consequently,

r σ + 1 n E j , k B r ( p ) ( max { d N ̂ tr H , 0 } ) σ C r σ + 1 n ( γ λ k ) σ i { 0 , 1 , , q } { j , k } | E j , k { M λ k 1 u i 0 } B r ( p ) | + C r σ + 1 n | E j , k B r ( p ) | C γ σ ( λ k λ 0 1 ) σ 1 ( λ 0 r ) σ 2 + C λ k 1 r σ 1 C γ σ q ( σ 1 ) + C λ k 1 ,

where 𝐶 is independent of 𝛾 and λ 0 . This completes the proof of Proposition 5.8. ∎

Proposition 5.9

Fix an exponent σ ( 1 , q q 1 ) . If λ 0 is sufficiently large (depending on 𝛾), then

r σ + 1 n G i , j , k B r ( p ) ( max { d N ̂ tr H , 0 } ) σ C γ σ q ( σ 1 )

for all 0 i < j < k q and all 0 < r 1 . Here, 𝐶 is independent of 𝛾 and λ 0 .

Proof

If λ 0 is sufficiently large (depending on 𝛾), then Corollary 5.7 implies that max { d N ̂ tr H , 0 } C γ λ k on the set G i , j , k , where 𝐶 is independent of 𝛾 and λ 0 . We distinguish two cases.

Case 1.  Suppose that 0 < r λ 0 1 . Since

G i , j , k Ω { u ̂ k = 0 } { 2 λ k 1 u ̂ k 1 0 } { 2 λ k 1 u k 0 } ,

Lemma 2.12 implies that | G i , j , k B r ( p ) | C λ k 1 r n 2 , where 𝐶 is independent of 𝛾 and λ 0 . Consequently,

r σ + 1 n G i , j , k B r ( p ) ( max { d N ̂ tr H , 0 } ) σ C r σ + 1 n ( γ λ k ) σ | G i , j , k B r ( p ) | C γ σ ( λ k r ) σ 1 C γ σ q ( σ 1 ) ,

where 𝐶 is independent of 𝛾 and λ 0 .

Case 2.  Suppose that λ 0 1 r 1 . Since

G i , j , k Ω { u ̂ k = 0 } { 2 λ k 1 u ̂ k 1 0 } { 2 λ k 1 u k 0 } { 4 λ 0 1 u j 0 } { 6 λ 0 1 u i 0 } ,

Lemma 2.13 implies that | G i , j , k B r ( p ) | C λ k 1 λ 0 1 r n 3 , where 𝐶 is independent of 𝛾 and λ 0 . Consequently,

r σ + 1 n G i , j , k B r ( p ) ( max { d N ̂ tr H , 0 } ) σ C r σ + 1 n ( γ λ k ) σ | G i , j , k B r ( p ) | C γ σ ( λ k λ 0 1 ) σ 1 ( λ 0 r ) σ 2 C γ σ q ( σ 1 ) ,

where 𝐶 is independent of 𝛾 and λ 0 . This completes the proof of Proposition 5.9. ∎

Corollary 5.10

Fix an exponent σ ( 1 , q q 1 ) . Suppose γ ( 0 , M 1 ) , where M > 2 denotes the constant in Corollary 5.5. If λ 0 is sufficiently large (depending on 𝛾), then

r σ + 1 n Σ ̂ B r ( p ) ( max { d N ̂ tr H , 0 } ) σ C γ σ q ( σ 1 )

for all 0 < r 1 . Here, 𝐶 is independent of 𝛾 and λ 0 .

Proof

This follows by combining Propositions 2.17, 5.2, 5.8, and 5.9. ∎

6 Proof of Theorem 1.3

In this section, we give the proof of Theorem 1.3. It suffices to consider the odd-dimensional case. (The even-dimensional case can be reduced to the odd-dimensional case by passing to the Cartesian product Ω × [ 1 , 1 ] R n + 1 .) Suppose that n 3 is an odd integer, and Ω = m = 0 q { u m 0 } is a compact, convex polytope in R n with non-empty interior. Let 𝑔 be a Riemannian metric on R n . We assume that the assumptions of Theorem 1.3 are satisfied. In other words, Assumptions 1.1, 1.2, 2.1, 4.1, and 5.1 are satisfied, and 𝑔 has nonnegative scalar curvature at each point in Ω.

Let us fix an exponent σ ( 1 , q q 1 ) . The results in the previous sections imply that we can find a sequence of domains Ω ̂ ( l ) and a sequence of smooth maps N ̂ ( l ) : Ω ̂ ( l ) S n 1 with the following properties.

  • For each 𝑙, Ω ̂ ( l ) is a compact, convex domain in R n with smooth boundary Ω ̂ ( l ) = Σ ̂ ( l ) .

  • For each 𝑙, m = 0 q { u m l 1 } Ω ̂ ( l ) Ω .

  • For each 𝑙, the map N ̂ ( l ) : Σ ̂ ( l ) S n 1 is homotopic to the Gauss map of Σ ̂ ( l ) .

  • We have

    (6.1) sup p R n sup 0 < r 1 r σ + 1 n Σ ̂ ( l ) B r ( p ) ( max { d N ̂ ( l ) tr H , 0 } ) σ 0

    as l .

Let us fix a Euclidean ball 𝑈 with the property that the closure of 𝑈 is contained in the interior of Ω. Note that U Ω ̂ ( l ) if 𝑙 is sufficiently large. In the following, we will always assume that 𝑙 is chosen sufficiently large so that U Ω ̂ ( l ) .

Proposition 6.1

There exists a uniform constant 𝐶 (independent of 𝑙) such that

Ω ̂ ( l ) F 2 d vol g C Ω ̂ ( l ) | F | 2 d vol g + C U F 2 d vol g

for every smooth function F : Ω ̂ ( l ) R .

Proof

The hypersurface Σ ̂ ( l ) = Ω ̂ ( l ) can be written as a radial graph with bounded slope. From this, it is easy to see that Ω ̂ ( l ) is bi-Lipschitz equivalent to the Euclidean unit ball, with constants that are independent of 𝑙. The assertion follows now from the corresponding estimate on the unit ball. ∎

Proposition 6.2

There exists a uniform constant 𝐶 (independent of 𝑙) such that

Σ ̂ ( l ) F 2 d σ g C Ω ̂ ( l ) | F | 2 d vol g + C Ω ̂ ( l ) F 2 d vol g

for every smooth function F : Ω ̂ ( l ) R .

Proof

The hypersurface Σ ̂ ( l ) = Ω ̂ ( l ) can be written as a radial graph with bounded slope. From this, it is easy to see that Ω ̂ ( l ) is bi-Lipschitz equivalent to the Euclidean unit ball, with constants that are independent of 𝑙. The assertion follows now from the Sobolev trace theorem on the unit ball. ∎

Proposition 6.3

We have

Σ ̂ ( l ) max { d N ̂ ( l ) tr H , 0 } F 2 d σ g o ( 1 ) Ω ̂ ( l ) | F | 2 d vol g + o ( 1 ) Σ ̂ ( l ) F 2 d σ g

for every smooth function F : Ω ̂ ( l ) R .

Proof

The hypersurface Σ ̂ ( l ) = Ω ̂ ( l ) can be written as a radial graph with bounded slope. From this, it is easy to see that Ω ̂ ( l ) is bi-Lipschitz equivalent to the Euclidean unit ball, with constants that are independent of 𝑙. Therefore, the assertion follows from (6.1) together with an estimate of Fefferman and Phong (see [1, Appendix A] and [2]). This completes the proof of Proposition 6.3. ∎

As in [1], we consider a boundary value problem for the Dirac operator. Let m = 2 [ n 2 ] . Let 𝒮 denote the spinor bundle over Ω. Note that 𝒮 is a complex vector bundle of rank 𝑚 equipped with a Hermitian inner product. We define a complex vector bundle ℰ over Ω by

E = S S m times .

Note that ℰ has rank m 2 .

As in [1], we may use the map N ̂ ( l ) : Σ ̂ ( l ) S n 1 to define a boundary chirality

χ ( l ) : E | Σ ̂ ( l ) E | Σ ̂ ( l ) .

Definition 6.4

Definition 6.4 (cf. [1, Definition 2.2])

Let { E 1 , , E n } denote the standard basis of R n . Let ω 1 , , ω n End ( C m ) be defined as in [1]. We define a bundle map

χ ( l ) : E | Σ ( l ) E | Σ ( l )

by

( χ ( l ) s ) α = a = 1 n β = 1 m N ̂ ( l ) , E a ω a α β ν s β

for α = 1 , , m .

Recall that the map

N ̂ ( l ) : Σ ̂ ( l ) S n 1

is homotopic to the Gauss map of Σ ̂ ( l ) . By [1, Proposition 2.15], we can find an 𝑚-tuple of spinors s ( l ) = ( s 1 ( l ) , , s m ( l ) ) defined on Ω ̂ ( l ) with the following properties.

  • s ( l ) solves the Dirac equation at each point in Ω ̂ ( l ) .

  • χ ( l ) s ( l ) = s ( l ) at each point on Σ ̂ ( l ) .

  • s ( l ) does not vanish identically.

Standard unique continuation arguments imply that

U | s ( l ) | 2 d vol g > 0

if 𝑙 is sufficiently large. By scaling, we may arrange that

(6.2) U | s ( l ) | 2 d vol g = m vol g ( U )

if 𝑙 is sufficiently large.

Combining Propositions 6.1, 6.2, and 6.3, we conclude that

(6.3) Σ ̂ ( l ) max { d N ̂ ( l ) tr H , 0 } F 2 d σ g o ( 1 ) Ω ̂ ( l ) | F | 2 d vol g + o ( 1 ) U F 2 d vol g

for every smooth function F : Ω ̂ ( l ) R . For each 𝑙, we apply (6.3) with F = ( δ 2 + | s ( l ) | 2 ) 1 2 , and send δ 0 . This gives

(6.4) 1 2 Σ ̂ ( l ) max { d N ̂ ( l ) tr H , 0 } | s ( l ) | 2 d σ g o ( 1 ) Ω ̂ ( l ) | s ( l ) | 2 d vol g + o ( 1 ) U | s ( l ) | 2 d vol g .

On the other hand, [1, Proposition 2.9] implies

(6.5) Ω ̂ ( l ) | s ( l ) | 2 d vol g + 1 4 Ω ̂ ( l ) R | s ( l ) | 2 d vol g 1 2 Σ ̂ ( l ) ( d N ̂ ( l ) tr H ) | s ( l ) | 2 d σ g 1 2 Σ ̂ ( l ) max { d N ̂ ( l ) tr H , 0 } | s ( l ) | 2 d σ g .

By assumption, the scalar curvature is nonnegative. Combining (6.2), (6.4), and (6.5), we conclude that

Ω ̂ ( l ) | s ( l ) | 2 d vol g 0

as l . Consequently, the sequence s ( l ) = ( s 1 ( l ) , , s m ( l ) ) converges in C loc ( Ω Ω , E ) to an 𝑚-tuple of parallel spinors s = ( s 1 , , s m ) which is defined on the interior of Ω. In particular, we can find a fixed matrix z End ( C m ) such that s α , s β = z α β at each point in the interior of Ω. Arguing as in [1, Section 4], we can show that 𝑧 is a scalar multiple of the identity matrix. Using (6.2), we conclude that 𝑧 is the identity matrix.

To summarize, s = ( s 1 , , s m ) is a collection of parallel spinors which are defined at each point in the interior of Ω and are orthonormal at each point in the interior of Ω. This implies that the Riemann curvature tensor of 𝑔 vanishes identically.

Since 𝑠 is parallel, 𝑠 can be extended continuously to Ω. We next consider the behavior of 𝑠 along the boundary faces. To that end, we recall the following result from [1].

Lemma 6.5

Lemma 6.5 (cf. [1, Lemma 3.2])

Let j { 0 , 1 , , q } . Then the set

{ u j = 0 } i { 0 , 1 , , q } { j } { u i < 0 }

is non-empty. Moreover, this set is a dense subset of Ω { u j = 0 } .

Lemma 6.5 follows from Assumption 1.1 together with the assumption that Ω has non-empty interior. We refer to [1] for a detailed proof.

Let us fix an arbitrary integer j { 0 , 1 , , q } . Arguing as in [1], we conclude that

ν j s α = a = 1 n β = 1 m N j , E a ω a α β s β

at each point in { u j = 0 } i { 0 , 1 , , q } { j } { u i < 0 } . In view of Lemma 6.5, the identity

ν j s α = a = 1 n β = 1 m N j , E a ω a α β s β

holds at each point in Ω { u j = 0 } . Therefore, if 0 j < k q and 𝑥 is a point in

Ω { u j = 0 } { u k = 0 } ,

then

m ν j , ν k = α = 1 m ν j , ν k s α , s α = 1 2 α = 1 m ν j s α , ν k s α + 1 2 α = 1 m ν k s α , ν j s α = 1 2 a , b = 1 n α , β , γ = 1 m N j , E a N k , E b ω a α β ω b α γ ̄ s β , s γ + 1 2 a , b = 1 n α , β , γ = 1 m N j , E a N k , E b ω a α β ̄ ω b α γ s γ , s β = m N j , N k

at the point 𝑥. In the last step, we have used the fact that ω a ω b + ω b ω a = 2 δ a b id for all a , b { 1 , , n } .

Finally, let us fix an integer j { 0 , 1 , , q } . The arguments in [1, Section 4] show that the second fundamental form of the hypersurface { u j = 0 } vanishes at each point in { u j = 0 } i { 0 , 1 , , q } { j } { u i < 0 } . Using Lemma 6.5, we conclude that the second fundamental form of the hypersurface { u j = 0 } vanishes at each point in Ω { u j = 0 } . This completes the proof of Theorem 1.3.

Award Identifier / Grant number: DMS-2103573

Award Identifier / Grant number: DMS-2403981

Funding statement: The authors were supported by the National Science Foundation under grants DMS-2103573 and DMS-2403981 and by the Simons Foundation.

A An estimate for the area of a level set of a convex function

Proposition A.1

Let m 2 be an integer, let f : R m R be a convex function, and let Σ : = { f = 0 } { f 0 } . Then Σ is a smooth hypersurface in R m . Moreover,

| Σ B r ( p ) | C ( m ) r m 1 .

Proof

Let us fix a unit vector ω S m 1 . Let Σ ω : = { f = 0 } { ω , f > 0 } and let V ω : = { y R m : ω , y p = 0 } denote the hyperplane orthogonal to 𝜔 passing through 𝑝. We denote by π ω : Σ ω V ω the orthogonal projection from the hypersurface Σ ω to the hyperplane V ω . Since 𝑓 is a convex function, the map π ω is injective. At each point on Σ ω , the absolute value of the Jacobian determinant of π ω is given by | det D π ω | = ω , f / | f | . Moreover, the image π ω ( Σ ω B r ( p ) ) is contained in V ω B r ( p ) . Thus, we conclude that

Σ ω B r ( p ) ω , f | f | | V ω B r ( p ) | C ( m ) r m 1 .

Finally, we integrate over ω S m 1 . This gives | Σ B r ( p ) | C ( m ) r m 1 . ∎

B Some elementary inequalities

Lemma B.1

The function t t cos t sin t is monotone decreasing for t ( 0 , π ) .

Proof

Note that t sin t cos t for all t > 0 . This implies

d d t ( t cos t sin t ) = t sin t cos t sin 2 t 0

for all t ( 0 , π ) . This completes the proof of Lemma B.1. ∎

Proposition B.2

Let α ( 0 , π 2 ) and β [ 0 , α ] . Then

sin ( 2 β ) sin ( 2 α ) sin ( 2 θ β ) sin ( 2 θ α ) 0

for all θ ( 0 , 1 ) .

Proof

It suffices to prove the assertion for β ( 0 , α ] . Using Lemma B.1, we obtain

2 θ α cos ( 2 θ α ) sin ( 2 θ α ) 2 θ β cos ( 2 θ β ) sin ( 2 θ β ) 0

for all θ ( 0 , 1 ) . This implies

d d θ ( sin ( 2 θ β ) sin ( 2 θ α ) ) = sin ( 2 θ β ) sin ( 2 θ α ) ( 2 α cos ( 2 θ α ) sin ( 2 θ α ) 2 β cos ( 2 θ β ) sin ( 2 θ β ) ) 0

for all θ ( 0 , 1 ) . From this, the assertion follows. ∎

Proposition B.3

Let α ( 0 , π 2 ) and β [ 0 , α ] . Then

| sin ( 2 β ) sin ( 2 α ) sin ( 2 θ β ) sin ( 2 θ α ) | 4 ( θ 1 ) α sin ( 2 α ) sin ( 2 θ α )

for all θ [ 1 , π 2 α ) .

Proof

Let θ [ 1 , π 2 α ) . Then

| sin ( 2 α ) sin ( 2 θ α ) | 2 ( θ 1 ) α , | sin ( 2 β ) sin ( 2 θ β ) | 2 ( θ 1 ) β .

This implies

| sin ( 2 θ α ) sin ( 2 β ) sin ( 2 α ) sin ( 2 θ β ) | 2 ( θ 1 ) ( α + β ) 4 ( θ 1 ) α .

From this, the assertion follows. ∎

Acknowledgements

The authors acknowledges the hospitality of Tübingen University, where part of this work was carried out.

References

[1] S. Brendle, Scalar curvature rigidity of convex polytopes, Invent. Math. 235 (2024), no. 2, 669–708. 10.1007/s00222-023-01229-xSearch in Google Scholar

[2] C. Fefferman and D. H. Phong, Lower bounds for Schrödinger equations, Conference on partial differential equations (Saint Jean de Monts 1982), Société Mathématique de France, Paris (1982), 1–7. 10.5802/jedp.249Search in Google Scholar

[3] M. Gromov, Dirac and Plateau billiards in domains with corners, Cent. Eur. J. Math. 12 (2014), no. 8, 1109–1156. 10.2478/s11533-013-0399-1Search in Google Scholar

[4] M. Gromov, Four lectures on scalar curvature, Perspectives in scalar curvature. Vol. 1, World Scientific, Hackensack (2023), 1–514. 10.1142/9789811273223_0001Search in Google Scholar

[5] M. Gromov, Convex polytopes, dihedral angles, mean curvature and scalar curvature, Discrete Comput. Geom. 72 (2024), no. 2, 849–875. 10.1007/s00454-024-00657-7Search in Google Scholar

[6] C. Li, A polyhedron comparison theorem for 3-manifolds with positive scalar curvature, Invent. Math. 219 (2020), no. 1, 1–37. 10.1007/s00222-019-00895-0Search in Google Scholar

[7] C. Li, The dihedral rigidity conjecture for 𝑛-prisms, J. Differential Geom. 126 (2024), no. 1, 329–361. 10.4310/jdg/1707767340Search in Google Scholar

[8] J. Wang, Z. Xie and G. Yu, On Gromov’s dihedral extremality and rigidity conjectures, preprint (2021), https://arxiv.org/abs/2112.01510. Search in Google Scholar

Received: 2023-12-10
Revised: 2025-06-18
Published Online: 2025-07-31

© 2025 the author(s), published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 29.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/crelle-2025-0048/html
Scroll to top button